163
Molecular cloning, characterization and relevance for microcystin-kinetics of Organic anion transporting polypeptides in model and native fish species Dissertation zur Erlangung des akademischen Grades eines Doktors der Naturwissenschaften an der Mathematisch-Naturwissenschaftliche Sektion Fachbereich Biologie vorgelegt von Konstanze Steiner Tag der mündlichen Prüfung: 7.11.2014 1. Referent: Prof. Daniel Dietrich 2. Referent: Prof. Bruno Hagenbuch Konstanzer Online-Publikations-System (KOPS) URL: http://nbn-resolving.de/urn:nbn:de:bsz:352-0-301748

Molecular cloning, characterization and relevance for

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Molecular cloning, characterization and relevance for

Molecular cloning, characterization and relevance for

microcystin-kinetics of Organic anion transporting polypeptides in model and native fish species

Dissertation zur Erlangung des akademischen Grades eines Doktors der Naturwissenschaften

an der

Mathematisch-Naturwissenschaftliche Sektion

Fachbereich Biologie

vorgelegt von Konstanze Steiner

Tag der mündlichen Prüfung: 7.11.2014 1. Referent: Prof. Daniel Dietrich

2. Referent: Prof. Bruno Hagenbuch

Konstanzer Online-Publikations-System (KOPS) URL: http://nbn-resolving.de/urn:nbn:de:bsz:352-0-301748

Page 2: Molecular cloning, characterization and relevance for
Page 3: Molecular cloning, characterization and relevance for

-Astor Piazzolla- (Invierno Porteño)

Page 4: Molecular cloning, characterization and relevance for

TABLE OF CONTENTS

GENERAL INTRODUCTION ................................................................................................................................... - 1 -

CYANOBACTERIA AND THEIR IMPACT ON THE ENVIRONMENT .................................................................................. - 3 -

MICROCYSTIN (MC) ..................................................................................................................................................... - 5 -

TOXICOKINETICS AND DYNAMICS OF MICROCYSTIN IN FISH ..................................................................................... - 7 -

DYNAMICS – RESPONSES TO MCS ............................................................................................................................................... - 7 -

LD50 and pathological alterations in organs and general physiological condition .............................................. - 7 -

Molecular mechanisms of MC ......................................................................................................................................................... - 8 -

KINETICS ....................................................................................................................................................................................... - 10 -

Absorption ............................................................................................................................................................................................ - 10 -

Distribution .......................................................................................................................................................................................... - 11 -

Metabolism .......................................................................................................................................................................................... - 12 -

Excretion ............................................................................................................................................................................................... - 13 -

TOXICOKINETICS AND –DYNAMICS OF OTHER MC CONGENERS IN FISH ............................................................................ - 14 -

ORGANIC ANION TRANSPORTING POLYPEPTIDES (OATPS) ................................................................................... - 15 -

CLASSIFICATION AND PHYLOGENESIS ...................................................................................................................................... - 15 -

GENERAL FEATURES ................................................................................................................................................................... - 16 -

TRANSPORTED SUBSTRATES ..................................................................................................................................................... - 16 -

TISSUE DISTRIBUTION ................................................................................................................................................................ - 18 -

TRANSPORT MODE AND DRIVING FORCE ................................................................................................................................. - 18 -

OATPS IN FISH .............................................................................................................................................................................. - 19 -

GOAL OF THIS STUDY ................................................................................................................................................. - 20 -

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER

OATP ......................................................................................................................................................................... - 21 -

ABSTRACT ................................................................................................................................................................... - 21 -

INTRODUCTION........................................................................................................................................................... - 23 -

METHODS .................................................................................................................................................................... - 25 -

RESULTS ...................................................................................................................................................................... - 30 -

DISCUSSION ................................................................................................................................................................. - 36 -

FUNDING ..................................................................................................................................................................... - 39 -

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON

CARP ......................................................................................................................................................................... - 41 -

Page 5: Molecular cloning, characterization and relevance for

ABSTRACT ................................................................................................................................................................... - 41 -

INTRODUCTION........................................................................................................................................................... - 43 -

MATERIAL AND METHODS ........................................................................................................................................ - 46 -

RESULTS ...................................................................................................................................................................... - 49 -

DISCUSSION ................................................................................................................................................................. - 55 -

ACKNOWLEDGEMENTS ............................................................................................................................................... - 58 -

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS .................................. - 59 -

ABSTRACT ................................................................................................................................................................... - 59 -

INTRODUCTION........................................................................................................................................................... - 61 -

MATERIAL AND METHODS ........................................................................................................................................ - 63 -

RESULTS ...................................................................................................................................................................... - 67 -

DISCUSSION ................................................................................................................................................................. - 76 -

CONCLUSIONS ............................................................................................................................................................. - 78 -

ACKNOWLEDGEMENTS ............................................................................................................................................... - 78 -

GENERAL DISCUSSION ........................................................................................................................................ - 79 -

EVOLUTIONARY RELATIONSHIP OF FISH AND MAMMALIAN OATPS ........................................................................ - 81 -

SUBSTRATE SPECIFICITY AND ORGAN SPECIFIC EXPRESSION OF OATPS ................................................................ - 84 -

IMPACT OF MC ACCUMULATION AND OATP EXPRESSION IN FISH: RELEVANCE FOR FISH .................................... - 89 -

RELEVANCE FOR THE INDIVIDUAL ............................................................................................................................................ - 89 -

RELEVANCE FOR THE ECOLOGICAL BALANCE ......................................................................................................................... - 93 -

IMPACT OF MC ACCUMULATION AND OATPS IN FISH: RELEVANCE FOR HUMANS ................................................. - 96 -

FINAL CONCLUSIONS ................................................................................................................................................... - 99 -

SUPPLEMENTAL MATERIAL ........................................................................................................................... - 101 -

SUPPLEMENTAL FIGURES ....................................................................................................................................... - 103 -

SUPPLEMENTAL TABLES ......................................................................................................................................... - 115 -

NOVEL SEQUENCES IDENTIFIED IN THE PRESENT STUDY ..................................................................................... - 121 -

MISCELLANEOUS .................................................................................................................................................... 135

LIST OF FIGURES .......................................................................................................................................................... 137

LIST OF TABLES ....................................................................................................................................................... - 138 -

ABBREVIATIONS ...................................................................................................................................................... - 139 -

Page 6: Molecular cloning, characterization and relevance for

DECLARATION OF SELF-CONTRIBUTION ................................................................................................................ - 141 -

ACKNOWLEDGEMENTS ............................................................................................................................................ - 143 -

SUMMARY ................................................................................................................................................................ - 145 -

ZUSAMMENFASSUNG ............................................................................................................................................... - 149 -

REFERENCES ........................................................................................................................................................ - 153 -

Page 7: Molecular cloning, characterization and relevance for

CHAPTER 1

GENERAL INTRODUCTION

Page 8: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Cyanobacteria and their impact on the environment

- 3 -

CYANOBACTERIA AND THEIR IMPACT ON THE ENVIRONMENT With one thousand million tons wet biomass and the oldest fossils found about 3.5 billion years ago,

Cyanobacteria (or blue-green algae) are one of the most abundant and oldest organisms on earth

(Schopf and Packer, 1987). Therefore, it is not surprising that they have a huge impact on various

ecological systems, including aquatic and terrestrial habitats. As they were the first phototrophic

microorganisms in history, they most likely played a major role in the oxygen accumulation in the

atmosphere and they still are – together with higher plants and algae - responsible for producing

organic biomass and are therefore important primary producers, building the basis of the aquatic food

web (Canfield, 2005; Sergeev et al., 2002). For humans they are additionally interesting as nitrogen

fixing species which enhance fertility of soil and agricultural wetlands (Whitton, 2000). Due to their

metabolites they are interesting for the development of pharmaceuticals like antibodies and anti-

cancer drugs as well as for herbicides and insecticides, moreover they are used as food supplement in

human and animal food (Falconer et al., 1999; Singh et al., 2005). Furthermore they are applicable for

alternative energies(Quintana et al., 2011).

However, as useful and important cyanobacteria were and still are, there is also a downside to their

existence. When biotic and abiotic factors are favorable for growing, they occur in masses, leading not

only to an unpleasant odor and water coloration but more importantly to eutrophication of water

bodies. Since this is accompanied with oxygen deficiency, decrease in species diversity, building of

toxic byproducts and consequently death of various animals including fish, they have tremendous

effects on the aquatic ecosystem. The fact that several metabolites produced by cyanobacteria are

highly toxic to aquatic and terrestrial species, adds to the problem. Since cyanobacteria can occur in all

kinds of water bodies, this can affect lakes and rivers used for recreational (and fishing) purposes as

well as aquaculture and drinking water reservoirs, which often leads to closure of water supplies. Only

recently a bloom in Lake Erie in the US caused the shutdown of water supplies for Toledo (Ohio) and

the adjacent countries, leaving 500,000 people without tap water (Gallo, 2014). Toxin contents in the

mentioned water bodies and their organisms make an impact on health, economy and ecology;

consequently there is a rising interest of scientist but also the public to understand cyanobacteria.

Main interests are particularly building of mass occurrences, the accompanied toxin production and -

release as well as the mechanisms of toxin uptake and incorporation in aquatic and terrestrial species.

Cyanobacteria have been classified as algae for more than 180 years, hence the historically given

synonym blue-green algae. Although they share characteristics with algae, including their size and the

blue-green and green pigments which enable them to produce oxygen via photosynthesis, they have

been taxonomically characterized as bacteria because they have some important characteristics of

prokaryotes such as the lack of a nucleus and ribosomes consisting of a 70s subunit (Mur et al., 1999).

Page 9: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Cyanobacteria and their impact on the environment

- 4 -

However, compared to other bacteria, the morphology of the about 2000 existing species is diverse,

including solitary cells, unicellular colonies, various filamentous species or even multicellular

organizations (Mur et al., 1999). Their diversity as well as physiological and ecological adaptions is an

advantage for their survival. Amongst these, the additional pigments phycocyanin and allophycocyanin

allow them to use light in deep water layers, which contributes to their benefit over other organisms

as it is inaccessible to the majority of green algae and plants (Mur et al., 1999; Oliver and Ganf, 2000;

van den Hoek et al., 1995). The ability to regulate their buoyancy by forming hollow gas vacuoles

promotes migration into these favored positions in the water column. Some species are able to survive

in low nutrient environments as they are able to fixate nitrogen in specialized heterocysts or even in

the same cell. Storage bodies and the formation of spores facilitate survival of long unfavorable

conditions (Oliver and Ganf, 2000). Hence, their worldwide occurrence in fresh and brackish water is

not surprising. Due to their numerous adaptions and their high toleration of several abiotic factors,

including high temperature, alkaline conditions and a high salt content in the water, they moreover

occur in extreme habitats like the dead sea with very high salt concentrations, hot springs, alkaline

water, cold glacial and polar lakes (Jungblut et al., 2005; Krienitz et al., 2003; Oren, 2000; Vincent,

2000; Ward and Castenholz, 2000).

These advantages as well as their quick asexual reproduction, allows them to proliferate quickly.

Favorable conditions including phosphor enrichment can therefore lead to mass occurrence (blooms).

Hence, bloom formation is often observed in stable, warm and nutrient rich lakes rather than in alpine

or Polar regions. These favorable conditions might be the result of natural factors but is more

commonly caused by anthropogenic influences like agricultural runoff or inadequate sewage

treatment (Bartram et al., 1999). Because of their buoyancy, blooms can appear very sudden and built

scums at the surface of the water. Depending on wind conditions they can accumulate especially in

bays or shorelines. The consequence of an established bloom is a higher biomass production and

therefore further eutrophication of the lake, including oxygen depletion and production of methane,

hydrogen sulfide and ammoniac, which are toxic for fish and other organisms in the lake. Since 46 of

the 2000 species built products which are toxic to aquatic and terrestrial animals, cyanobacterial

blooms can pose a severe risk. Especially the senescence of a bloom after the lifespan of the cells can

lead to a sudden release of high concentrations of toxins and was often reinforced anthropogenically

by water treatment (e.g. chlorination or algicide application, (Hrudey et al., 1999; Sivonen and Jones,

1999). In fact, it was estimated that up to 75 % of bloom sample material contained toxins (Sivonen

and Jones, 1999). Since a bloom is often not dominated by one species but rather a mixture of toxic and

non-toxic strains ,as well as the fact that several species produce different kind of toxins, a

simultaneous exposure to various toxins is possible (Funari and Testai, 2008).

Page 10: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Microcystin (MC)

- 5 -

Numerous cyanotoxins have been described. In general they are named after the organism they have

been first isolated from but were mostly found to be produced by several other species as well.

Cyanotoxins can be divided according to their target or their structure. Lipopolysachharides, which

are a common part of the bacterial outer cell membrane of gram-negative prokaryotes, are mainly

toxic in that matter that they are elicit irritants and have allergic effects due to the fatty acid

component (Sivonen and Jones, 1999; Weckesser et al., 1979). Some alkaloids like saxitoxin and

anatoxins mediate neurotoxicity, which has also been suggested for the non-proteinogenic amino acid

β-methylamino-L-alanine (BMAA) (Cox et al., 2003; Funari and Testai, 2008). A general cytotoxicity

has been observed for the alkaloid cylindrospermopsin, whereas the oligopeptides microcystin (MC)

and nodularin are hepatotoxic (Ohtani et al., 1992; Sivonen and Jones, 1999). Even though MC is

primary hepatotoxic, it is targeting other organs as well, and can therefore be considered cytotoxic.

Amongst these and other toxins, MC is in fresh and brackish water one of the most abundant and toxic

ones at the same time (Sivonen and Jones, 1999).

MICROCYSTIN (MC) MC was first discovered in Microcystis aroginosa, a bloom forming species in eutrophic lakes, but it has

later also been isolated from other species including Anabaena sp, Planktothrix sp, and Nostoc sp,

Anabaenopsis sp and Hapalosiphon sp. (Botes et al., 1982; Botes et al., 1984). The trigger for the MC

production as well as the biological role remains still unclear and several factors have been discussed

(Kaplan et al., 2011; Neilan et al., 2013).

MCs are oligopeptides containing 7 amino acids (FIGURE 1). Alterations in the amino acid composition

and their methylation state make them the most diverse group of cyanotoxins, with more than 100

structural variants (congeners). Therefore the molecular weight varies between 900 und 1100 Dalton

(Da). MC is composed of seven amino acids with the general structure cyclo-(D-Ala1-X2-D-MeAsp3-Z4-

Adda5-D-Glu6-Mdha7). This includes a D-alanine at position 1, a D-Glutamic at position 6 and two

variable L-amino acids at positions 2 and 4. Moreover most MCs contain the amino acid (2S,3S,8S,9S)-

3-amino-9-methoxy-2,6,8-trimethyl-10-phenyldeca-4,6-dienoic acid (Adda) at position 5, which is

restricted to cyanobacteria and crucial for MC activity (An and Carmichael, 1994). D-erythro-β-

methylaspartic acid (MeAsp) can be found at position 3 and a N-methyldehydroalanine (Mdha) at

position 7. Mainly structural variations arise by exchanging the 2 variable L-amino acids, which are

therefore used for nomenclature of MCs, e.g. MC-LR for the MC congener with leucine and arginine in

position 2 and 7, respectively.

Once released into the water, total concentrations of dissolved MC vary between 10 µg/L and 350 µg/L

in surface water and can reach up to 25000 µg/L in remaining parts of the water body (Sivonen and

Jones, 1999; van Apeldoorn et al., 2007). However it has to be kept in mind, that intracellular toxin

Page 11: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Microcystin (MC)

- 6 -

concentrations are generally much higher than MC concentrations dissolved in the water (van

Apeldoorn 2007). Additionally, MCs are extremely stable in their natural environment. The main break

down takes place via photolysis (photochemical degradation) and bacterial degradation (Maruyama et

al., 2003; Park et al., 2001; Rapala et al., 2005; Tsuji et al., 1994). Breakdown might only need 2

weeks for >90 %, but it can take longer than 6 weeks if conditions like temperature and structure of

the water body are unfavorable (Sivonen and Jones, 1999). Even though catalyzed breakdown via

bacterial and photolysis is for sure more sensitive, a huge impact on decreasing of the concentration

has most likely the dilution into the water body (Jones and Orr, 1994). Treatment of the toxin

containing water however is rather difficult, since MCs are resistant to boiling. Also chemical

hydrolyses does not work at neutral pH but only under 6 M HCl and high temperature (Harada, 1995).

Figure 1: Structure of Microcystin-LR. Shown is the general structure of the heptapeptide with the amino acids in different colors. The variable positions 2 and 4 are Leucine (L) and Arginine (R) in Microcystin-LR. Adda: (2S,3S,8S,9S)-3-amino-9-methoxy-2,6,8-trimethyl-10-phenyldeca-4,6-dienoic acid; Mdha: N-methyldehydroalanine and MeAsp: D-erythro-β-methylaspartic acid.

Terrestrial animals and humans can be exposed to the toxin through various routes. A major uptake

route is through ingestion, when the toxin containing waterbody is used as drinking water reservoir.

This has often been observed to lead to acute exposures resulting in gastroenteritis and liver damages.

Symptoms after oral exposure are diarrhea, vomiting, flu symptoms, skin rashes, mouth ulcers, fevers,

irritation of eye and occur within seven days of exposure. Another way of oral ingestion of minor

concentrations is the consumption of sea food and algal health food products, containing

cyanobacterial cells. When consumed regularly, a chronic exposure is conceivable and has been

associated with tumors and carcinomas in rodents and humans (Fujiki and Suganuma, 2011). Minor

Page 12: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 7 -

routes of exposure are inhalation (e.g. via showering or recreational use of the water body) or dermal

exposure which might lead to allergic reactions (Falconer et al., 1999).

Because of the large impact on human health, the World Health Organization (WHO) recommends

1 µg/L as the guideline value for the maximum MC-LR amount in drinking water (WHO, 1998). Based

on this, a limitation of 0.04 µg MC-LR/kg body weight (bw)/day was suggested for tolerable daily

intake (Kuiper-Goodman et al., 1999).

TOXICOKINETICS AND DYNAMICS OF MICROCYSTIN IN FISH Cyanobacterial blooms pose a threat not only to terrestrial animals and humans but also to aquatic

organism since their possibility of avoidance is limited. These organisms often have to deal with

ingestion of toxic cells on the one hand but also dissolved toxins in the water on the other hand. For

several reasons fish are of special interest within the aquatic livestock: Intoxication of fish might lead

to losses in aquaculture or even jeopardize human’s health as consumer since fish is an important food

source. From the ecological perspective differences in species sensitivity towards MC are likely to lead

to a shift in the species distribution in an ecological system. Indeed, mortality of fish belonging to all

trophic levels has been observed upon a cyanobacterial bloom all over the world (Jewel et al., 2003;

Rodger et al., 1994; Toranzo et al., 1990). This is likely to be the consequence of several factors

including toxin production but also other conditions like increasing ammonium levels or low oxygen,

which occurs during a bloom (Jewel et al., 2003; Kangur et al., 2005). This is especially supported by

the fact that natural occurring concentrations of MC, neither dissolved in the water nor in the

cyanobacterial cells are high enough to lead to a lethal effect alone (Ibelings and Havens, 2008). Even

though mass mortalities cannot exclusively be associated with MC, sublethal and chronic effects have

been shown in various field- and laboratory studies. Kinetics of MC in fish has been investigated

regarding uptake and elimination time of MC, organ distribution of the toxin and biotransformation.

Dynamics – Responses to MCs

In the last two to three decades, research mainly focused on the dynamics of MC in fish, thereby

looking at organ toxicology, changes in behavior and biochemical as well as molecular alterations.

LD50 and pathological alterations in organs and general physiological condition

Direct exposure of MC-LR to fish under controlled conditions demonstrated that the toxin is highly

toxic or even leads to mortality. Depending on the exposure route (gavaging or intraperitoneal

injection (ip)) of a single application between 25 µg/kg bw and 6600 µg/kg bw resulted in death of all

experimental fish within 5 h to 4 days. All studies investigating organ pathology in fish consistently

found alterations in the liver. Morphological changes in the liver included a discoloration, a spongy

appearance as well as an enlargement of the organ, which all indicate severe pathology (Ernst et al.,

Page 13: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 8 -

2006; Kotak et al., 1996; Råbergh et al., 1991). Damage of hepatocytes can be observed as early as 1 h

after gavaging including rounding of hepatocytes, condensed cytoplasm, loss of cell-cell contact as well

as apoptosis and necrosis of cells (Fischer and Dietrich, 2000). Therefore the liver seems to be the

primary target, as has been shown in humans and rodents as well. Even though effects on the liver

seem to be similar like in mammals, the latter rather die from hemorrhagic shock whereas fish die

from direct liver failure due to necrosis (Ibelings and Havens, 2008). In addition some investigations

also observed an impact on kidney and intestine. For the kidney, degenerative changes of the renal

proximal tubular cells included cell vacuolization and cell shredding (Ernst et al., 2007). In the

intestinal mucosa pyknotic nuclei and apoptotic cells were observed 12h post dosing (Fischer and

Dietrich, 2000). Interestingly, toxic effects vary in different species. Whereas development of

pathology in salmonids takes longer after MC exposure, toxic effects can be seen quite quickly and at

lower concentrations in common carp (Cyprinus carpio, Fischer and Dietrich, 2000; Fischer et al.,

2000). Moreover, in studies using common carp as experimental fish, pathology of the kidney and

gastrointestinal tract (GIT) was observed, whereas these pathologies could not be observed in

salmonids (Fischer and Dietrich, 2000; Kotak et al., 1996; Råbergh et al., 1991; Tencalla et al., 1994).

Ernst et al. (2006) observed pathological changes in the European whitefish (Coregonus lavaretus),

these however have only been seen sporadically and not in all fish.

Behavioral changes have been seen in a dose-dependent manner and included increased activity and

hyperventilation (Baganz et al., 2004; Baganz et al., 1998). Numerous studies observed changes in

swimming activity with increasing concentrations of MC (Carbis et al., 1996; Råbergh et al., 1991;

Tencalla et al., 1994), which has been associated with a possible neurotoxicology of MC-LR (Cazenave

et al., 2008).

On the physiological level a dose-dependent increase upon MC exposure of intracellular enzymes like

lactatedehydrogenase, alaninaminotransferase and aspartateaminotransferase has been

demonstrated (Malbrouck and Kestemont, 2006). Several additional alterations have been discussed,

including cardiovascular changes like increased heart rate and stroke volume, decrease of ion

homeostasis including sodium and chloride levels (Best et al., 2001; Bury et al., 1996; Gaete et al.,

1994). The latter has been associated with an inhibition of the Na+/K+-ATPase activity. The impact of

MC however is not definite since exposure experiments have mainly been conducted with crude

cyanobacterial extracts, whereas pure MC only had little or no effect, leading to the assumption that

rather additional compounds played a crucial role.

Molecular mechanisms of MC

On a cellular level, MC inhibits eukaryotic Threonin/Serin phosphatases, mainly Protein Phosphatase

(PP)1 and PP2A, but also PP4 and PP5 and PP2B (Hastie et al., 2005; Honkanen et al., 1990;

Page 14: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 9 -

Yoshizawa et al., 1990). The mechanism of inhibitions has been investigated with PP1 and PP2A and

mainly consists of a two-step mechanism, where MC first binds to the enzyme non-covalently within

minutes to inhibit it, while the covalent binding takes several hours. Thereby MC interacts with

different sites of the PPs: the non-covalent bond is built by the hydrophobic Adda side chain of MC at

position 5, which is binding in the hydrophobic groove of the PP, next to the active site. Additionally

the MeAsp residue at position 3 is blocking the access to the active center of the enzyme by interacting

with Arginine and Tyrosine residues of the PPs (Maynes et al., 2006). Moreover the d-glutamic acid at

position 6 plays a role in toxicity since its carboxyl group shares two water molecules with the

catalytic metal atoms. Finally the irreversible covalent bond is built mediated by the Mdha residue at

position 7 of MC (Craig et al., 1996). Even though most of the above described mechanisms have been

investigated with mammalian PPs, inhibition of PPs appear also to be the main target of MCs in fish,

since the activity of the enzyme is reduced in a time dependent manner upon MC exposure (Råbergh et

al., 1991; Tencalla and Dietrich, 1997). Inter species variations in that manner do not seem to differ

since the median inhibitory concentration (IC50) upon MC-LR induced PP inhibition was similar in carp

and in rainbow trout (Oncorhynchus mykiss) (Tencalla and Dietrich, 1997; Tencalla, 1995; Xu et al.,

2000).

Since PPs are responsible for dephosphorylating various proteins while most protein kinases at the

same time still perform phosphorylation of proteins, this inhibition consequently leads to

hyperphosphorylation of various proteins (Falconer and Yeung, 1992; Wickstrom et al., 1995). This

disruption of the cellular phosphorylation/dephosphorylation balance has manifold consequences

including alteration of the cytoskeleton with subsequent deformation of hepatocytes, loss of cell-cell

adhesion, disruption of cellular metabolism, interferences with signal transduction and disturbance of

cell cycle control (Craig and Holmes, 2000; Falconer and Yeung, 1992). All of this leads to necrosis,

apoptosis, oxidative stress, tumor production and genotoxicity. Since key regulators of several

pathways are affected from the PP inhibition, it is still very poorly understood which factors lead to

which outcome. In general, acute doses of MC seem to induce alterations in the cytoskeleton, oxidative

stress and apoptosis, whereas a chronic exposure more likely leads to cellular proliferation and tumor

promotion (Dietrich and Hoeger, 2005). As for the underlying mechanisms, the involvement of several

kinases and transcriptional factors have been suggested to be involved in genotoxicity, apopotosis,

alterations in cell-cycle control and production of reactive oxygen species (ROS), which itself leads to

further downstream outcomes (summarized in Campos and Vasconcelos, 2010). In fish, the formation

of ROS was determined by either measuring directly the ROS level or the activity of antioxidative

enzymes including superoxide dismutase (SOD), catalase (CAT), glutathione-peroxidase (GPx) and

glutathione-reductase (GR). Moreover, the amount of glutathione (GSH)/ Glutathion-Disulfid (GSSG)

was used as indicator for oxidative stress since ROS is reduced through an oxidation of GSH, resulting

in GSSG. This reaction is catalyzed by GPx, whereas GR is reducing GSSG to GSH again. Another

Page 15: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 10 -

approach was measuring the increase of products resulted from lipidperoxidation, which is a

biomarker for oxygen-mediated toxicity. An increase of ROS in carp hepatocytes could be observed

within 6 h of exposure, while at the same time the activity of ROS-eliminating enzymes including SOD,

CAT and GPx increased (Li et al., 2003). Similar was observed in loach (Misgurnus mizolepis) liver after

a long term exposure with MC containing cells. After 28 days the activity of antioxidant enzymes was

significantly increased (Li et al., 2005). The lipidperoxidation on the other hand remained stable,

which was also observed in zebrafish (Danio rerio), suggesting that oxidative stress was reduced by

the activation of these enzymes (Chen et al., 2012; Li et al., 2005). However some other studies

observed differences in the activation between the enzymes and the amount of lipidperoxidation,

which is likely due to different concentrations and exposure times of MC and the target organ

investigated (Chen et al., 2012; Pavagadhi et al., 2012; Prieto et al., 2006). This is supported since a

time- and concentration dependent activation of the antioxidant enzymes has been observed,

suggesting an increase in activity shortly after the exposure and decline in the course of time (Jiang et

al., 2011; Jiang et al., 2013; Li et al., 2010). At low MC concentrations, antioxidant enzymes are able to

eliminate oxidative stress, while at high concentrations oxidative stress is induced, resulting in lipid

peroxidation. Also the alteration of GSH-levels upon MC exposure revealed ambiguous results,

detecting either no difference (Adamovský et al., 2007; Amado et al., 2011) or a decrease of GSH. This

decrease was accompanied with an increase of GSSG, therefore suggesting a role of GSH in

detoxification processes (Chen et al., 2012; Jiang et al., 2011; Li et al., 2003). Depletion of GSH upon

MC exposure however cannot be ascribed to detoxification alone since it is involved in at least two

additional pathways involved in MC intoxication: On the one hand GSH is a substrate of the glutathione

peroxidase and involved in detoxification as described above. On the other hand GSH plays an

important role in the kinetics of MC, more specifically in the biotransformation and uptake, as

described later

Kinetics

The severity of MC intoxication mainly depends on its toxicokinetics. Important factors which have to

be taken into consideration are exposure route as well as transport into the blood system and

therefore bioavailability of the toxin. Moreover it is important to know which organs are likely to be a

target for the toxin and how quickly the toxin can be metabolized and excreted from the body again.

Absorption

Various routes of exposure are conceivable for fish. One way is the uptake of dissolved MC in the water

through the gills, intestine or integument. Another way is the digestion of cyanobacterial cells

containing MC. Ibelings and Havens (2008) suggested that feeding and subsequent uptake via the

digestive system is the most important route for ingestion since high concentrations of dissolved

Page 16: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 11 -

toxins are rather an exception due to mixing, photolysis and bacterial degradation. Because of that, fish

filtering food like herbivorous and planktivorous fish are more likely to be exposed to MC than

omnivorous and piscivorous fish. Nevertheless, toxins are also found in higher trophic levels,

suggesting an indirect uptake via toxin-contaminated food. Whether this transfer within the food chain

more likely leads to a biomagnification or a biodilution is still discussed as data are inconclusive (Chen

and Xie, 2005; Ibelings et al., 2005; Xie et al., 2005; Zhang et al., 2009b). Recently a quantitative meta-

analysis of 42 studies revealed that biodilution is the dominant process in aquatic foodwebs although

zooplanktivorous fish show some potential for biomagnification (Kozlowsky-Suzuki et al., 2012).

Notably, also fish belonging to related trophic groups exhibit a variation in their MC (Christoffersen,

1996; Deblois et al., 2008).

In laboratory experiments ip administration was more harmful than gavaging, as shown in carp and

rainbow trout (Carbis et al., 1996; Tencalla et al., 1994). Obviously the digestive system represents a

barrier to the toxin and minimizes its bioavailability in the blood and therefore for also the availability

for other organs.

Distribution

In fish, field studies as well as laboratory studies demonstrated the first detection of MC after exposure

in gut contents followed by the liver and feces, additionally some amounts have also been found in

muscle tissue (summarized in Malbrouck and Kestemont, 2006). In the fish intestine enzymatic

degeneration of MC occurs rather slowly, resulting in a high concentration of MC. To reach the

systemic blood however it needs to cross the epithelium of the ileum. Indeed it has been shown that

only less than 5 % of the MC orally exposed to rainbow trouts enters the blood stream (Tencalla and

Dietrich, 1997). Moreover it is assumed that absorption via the ileum depends on the anatomy of the

intestine. Rainbow trouts for example have a short and thick walled gut, whereas cyprinids possess a

longer intestine, therefore providing more surface for absorption (Fischer and Dietrich, 2000), due to

that it is likely that carp are more sensitive to MC. Indeed, oral exposure of freeze dried M. aeruginosa

revealed that rainbow trout died within 96 h at 6600 µg/kg bw MC-equivalent (equ.) while for carp an

exposure of less than 1700 µg/kg bw MCequ. was lethal (Tencalla, 1995; Tencalla et al., 1994).

After crossing the intestine, MC is bioavailable in the systemic blood. Highest accumulations of MC was

found in the liver of fish taken from MC containing lakes as well as in the liver of fish force fed with MC

or cyanobacteria. Thereby, the liver was not only the organ with the highest MC accumulations but

also the organ, where MC was mostly detected first (as soon as 1h after oral application). Smaller

amounts of MC and/or later occurring accumulations have been found in the kidney. Therefore the

main target is the liver, which is in accordance with the observed organ pathologies described above.

Additionally accumulations have been found in other organs including muscle and brain without

Page 17: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 12 -

necessarily resulting in pathologies (Fischer and Dietrich, 2000; Moutou et al., 2012; Singh and

Asthana, 2014; Williams et al., 1997; Williams et al., 1995).

MC is unlikely to just passively diffuse through the membrane due to its structure and size. Therefore

additional factors are limiting for the bioavailability of MC, which is a transport system. Likewise, only

those organs expressing the mentioned transporters can be targets for MC. This explains why only up

to 17 % of MC applied ip into salmon actually reached the different tissues (Williams et al., 1997;

Williams et al., 1995). Indeed, it has been shown that hepatocellular uptake of MC is inhibited by bile

salts and BSP, which suggested an uptake via a multispecific bile acid transport system (Runnegar et

al., 1995). Later, Fischer et al. (2005) were the first to proof in a Xenopus laevis expression system that

MC-LR is transported by human and rat Organic Anion transporting polypeptides (OATPs). An OATP-

mediated uptake was further suggested in the brain as murine whole brain cells and neurons

accumulate MC and at the time express OATPs at the RNA and protein level. Moreover, cytotoxicity

induced by MC was reduced when murine whole brain cells and neurons where co-incubated with

typical OATP substrates (Feurstein et al., 2009; Feurstein et al., 2010).

Metabolism

The recovery in fish from certain severe pathologies after an MC-exposure as well as the clearance of

MC suggests that either covalently bound MCs are metabolized in the course of time, or proteins and

peptides with covalently bound MC are replaced by de novo protein synthesis (Adamovský et al., 2007;

Fournie and Courtney, 2002; Malbrouck et al., 2003; Williams et al., 1995). It could be demonstrated

that major pathways for xenobiotic biotransformation into more polar conjugates also play a role in

the MC detoxification, therefore the water solubility increases and MC is likely to be less toxic and can

be easier excreted. The presence of glutathione and cysteine (CYS) conjugates after ip injection of MC-

LR could be identified for the first time in mouse and rat liver. Thereby the conjugated forms of MC

exhibited a reduction in toxicity compared to the native toxin MC-LR (Kondo et al., 1992; Kondo et al.,

1996). Later,Pflugmacher et al. (1998) concluded that the conjugation of MC-LR with GSH also seems

to be the first step in detoxification in aquatic macrophytes, invertebrates, fish and fish eggs as they

found a MC-LR/GSH conjugate formed by a soluble Glutathion-S-transferase (GST) 30 min after

incubation of an enzyme extract isolated from the aquatic organisms with MC-LR. Ito et al. (2002)

report a similar affinity of MC-LR and the two conjugated forms MC-LR/GSH and MC-LR/CYS to PP1C,

PP1y and PP2A, therefore suggesting that the lower toxicity of the conjugates is due to a carrier

mediated transport system. Only recently several studies used novel analytical methods to measure

the amount of MC and the conjugates quantitatively. Thereby the GSH conjugated form was

comparatively low or even not detectable whereas the CYS conjugated form was dominant (He et al.,

2012; Zhang et al., 2009a; Zhang et al., 2009b; Zhang et al., 2012). Consequently, Zhang et al. (2012)

suggested that CYS conjugation might play a more important role than GSH conjugation. However, Li et

Page 18: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 13 -

al. (2014a) demonstrated that injection of MC-RR/GSH leads to 97-fold increase of MC-RR/CYS in the

kidney confirming thereby assumptions, that MC/GSH is an intermediate metabolite which is rapidly

transformed into MC/CYS (He et al., 2012; Li et al., 2014a; Pflugmacher et al., 1998). Notably,

accumulation and excretion of conjugated MC took mainly place in the kidney as observed in various

investigations (He et al., 2012; Ito et al., 2002; Li et al., 2014a).

Additionally, indirect evidence supports the metabolism of MC through the GSH pathway. Numerous

studies investigated expression level and activity of the enzyme GST, which is catalyzing the

conjugation reaction with GSH. An increase of the GST mRNA expression level as well as its activity

suggested the involvement of GSH in detoxification of MC-LR (Adamovský et al., 2007; Chen et al.,

2012; Fu and Xie, 2006; Jiang et al., 2011; Sotton et al., 2012; Wang et al., 2006; Zhang et al., 2011).

However, a similar amount of investigations also found ambiguous data referring GST expression

levels or activity upon MC exposure or observed even a decrease in both (Amado et al., 2011; Li et al.,

2008; Malbrouck et al., 2003; Pavagadhi et al., 2012). The ambiguousness in this data is most likely

due to different factors used in these investigations, including fish species, MC-concentration and -

exposure time, investigated organ and so forth. Hence, drawing conclusions from this endpoint is

rather difficult.

As mentioned above, this is the second function of GSH in association with MC intoxication. It is

conceivable that conjugation of MC with GSH lead to a depletion of the GSH pool in the cell and

therefore to a diminished reduction of ROS.

Excretion

After a single application of MC, toxin amounts in the liver and intestine decreased after reaching their

maximum which further demonstrates excretion of the metabolized MC (Mohamed and Hussein, 2006;

Tencalla and Dietrich, 1997; Williams et al., 1997; Williams et al., 1995). Maximum concentrations of

MC were found 5 h after injection in the liver, which decreased about to half of the amount at 48 h.

Also after about 48 h the liver recovered as it turned from a pale yellow color to a more blood red hue

(Williams et al., 1997; Williams et al., 1995). Nevertheless, liver pathologies are still observed beyond

24 h, suggesting that parts of the conjugated MC gets excreted, but an accumulation in the liver still

takes place.

Also, fish chronically exposed to MC seem to eliminate most of the toxin, although this seems to be

species- and concentration-dependent as shown in various studies. While the carp hepatopancreas

was completely free again of MC after a month of exposure to Microcystis and an additional month in

clean water, the hepatopancreas of nile talipia (Oreochromis niloticus) exhibited only minor removal of

MC in similar exposure- and depuration-times. Another study could show in yellow catfish

(Pelteobagrus fulvidraco) that the amount of toxin in the diet is essential in the elimination of the toxin.

Page 19: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Toxicokinetics and dynamics of Microcystin in fish

- 14 -

It should also be noted that toxin and congener compositions in that studies probably varied

(Adamovský et al., 2007; Dong et al., 2012; Palikova et al., 2011).

The main way of excretion was suggested to take place via the bile, which is supported by the fact that

bile extracts from rainbow trout show an inhibition in their PP-activity after exposing them with

cyanobacterial cells (Sahin et al., 1996). In mice sublethal exposure of tritium labeled MC-LR revealed

an excretion of about 9 % of the applied toxin in the urine while about 14.5 % was excreted in the

feces (Robinson et al., 1991). This as well as the fact that major amounts of the conjugates have been

found in the kidney of fish as described above, leads to the assumption that renal excretion might also

play an additional role in fish as well. However, to our knowledge direct proof of that is still lacking.

Toxicokinetics and –dynamics of other MC congeners in fish

Most of the data described above has been conducted exposing fish to pure MC-LR or to cyanobacterial

cells containing a mixture of different congeners. Investigations with other congeners are minor in

comparison; however a few studies with MC-RR indicate that the general dynamics are comparable.

This includes the organotropism with highest amounts of toxin found in the liver and minor amounts

in kidney, brain and muscle (Cazenave et al., 2005; Chen et al., 2009; Xie et al., 2005). Also, changes in

behavior and alterations on a molecular level seem to be similar (Cazenave et al., 2006; Cazenave et

al., 2008; Lei et al., 2008; Pavagadhi et al., 2012; Prieto et al., 2006). The LD50 for MC-RR in zebrafish

is 4480 µg/kg bw, this however is difficult to compare with an LD50 for MC-LR since it has not been

investigated in zebrafish using the same application form (Zhao et al., 2012). The toxicity in mice

however is less severe as shown with an approximately 10 fold lower LD50 for MC-RR (Krishnamurthy

et al., 1986; Lovell et al., 1989; Stoner et al., 1989; Watanabe et al., 1987). This might be explained by

variations in the toxicodynamics as the IC50 for PP1 and PP2a were higher for MC-RR compared to MC-

LR, MC-LW and MC-LF, suggesting a lower inhibition with MC-RR (Monks et al., 2007). Other

investigations on the other side found that the inhibition of PPs is comparable between several

congeners including MC-LR, MC-RR, MC-LW and MC-LF and concluded that the less severe toxicity of

MC-RR is rather due to toxicokinetics than –dynamics (Fischer et al., 2010; Höger et al., 2007). Indeed,

Fischer et al. (2010) used OATP1B1/OATP1B3 transfected HEK293 cells to investigate differences in

the cytotoxicity induced by different MC-congeners. The authors found the highest cytotoxicity with

MC-LF and MC-LW, followed by MC-LR, whereas MC-RR induced less severe cytotoxicity. They

concluded that the differences in toxicity of various congeners are due to specific uptake properties for

each congener. Therefore the expression of specific OATP transporter plays an important role in the

selective uptake of MC-congeners.

Page 20: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Organic anion transporting polypeptides (OATPs)

- 15 -

ORGANIC ANION TRANSPORTING POLYPEPTIDES (OATPS) The fact that OATPs are involved in the uptake of MC and therefore play a major role in the uptake and

distribution of the toxin in mammalian organisms makes them interesting for understanding

toxicokinetics in fish as well. However, compared to mammals, only little is known about Oatps in fish

species. In the following, an overview will be given on some important characteristics of OATPs

investigated mainly in mammals. At the end, a summary will be given about the current knowledge of

Oatps in fish.

In humans, there are approximately 400 transporter-like genes described in the ATP binding cassette

(ABC) and solute carrier (SLC) superfamily, about 30 of that belong to the so called “drugtransporters”

(Keogh, 2011; Tweedie et al., 2013). Their primary function is the transport of nutrients and

endogenous substrates like sugar, amino acids, nucleotides and vitamins across cell membranes (You

and Morris, 2007). Additionally structural similar substances can be recognized. Since these

substances can also be xenobiotica like pharmaceutical drugs or toxins, these transporters play an

important role in the bioavailability and efficiency of a therapy based on drugs. Moreover it might

determine the dose related effects a toxin might have on the organism. OATPs are members of SLC

superfamily, therefore having an impact on the distribution of physiological substrates but also drugs

and toxins.

Classification and phylogenesis

Within the superfamily of solute carriers, OATPs/Oatps are members of the solute carrier organic

anion transporting family (SLCO). The first OATP was isolated 1994 from rat, within the next ten years

52 OATP members were identified. OATPs with more than 40 % amino acid identity belong to the

same family which is indicated with numbers (e.g. OATP1, OATP2). Numbering ensued according to

chronology. An identity of more than 60 % classifies a member within the same subfamily as indicated

with letters (e.g. OATP1A, OATP1B). Several individuals of one subfamily, which derived through gene

duplication within one species are differentiated with another number (e.g. OATP1A1, OATP1A2).

According to this classification the human and rodent Oatps cluster into 6 families and 13 subfamilies

(summarized in (Hagenbuch and Meier, 2004). However, with about 250 additional members being

identified in over 40 species in the last ten years it turned out that this given threshold of 40 and 60 %

is not absolute and can be less when OATP members of different species are compared, even though an

orthology is clearly given (summarized in (Hagenbuch and Stieger, 2013). In that case newly identified

members are named according to their orthologous. In general, the symbols for proteins and genes

depend on the nomenclature guideline of the species. For humans and rodents proteins are given in

capital letters (OATP1A1) whereas for fish only the first one is a capital letter (Oatp1a1). The gene

symbol is given in italics in humans, rodents and fish, but all letters are written in capitals for humans

Page 21: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Organic anion transporting polypeptides (OATPs)

- 16 -

(SLCO1A2), only the first one is written in capitals for rodents (Slco1a1) or none is written in capitals

in zebrafish (slco1a1).

General features

OATPs generally consist of 643-722 amino acids with a molecular mass between 80 and 90 kilo Dalton

(kDa) (summarized in Hagenbuch and Meier, 2004). Computational predictions suggest they consist of

12 transmembrane domains (TMD) and 6 extracellular as well as 5 intracellular loops (Figure 2,

summarized in Hagenbuch and Gui, 2008). Conserved structures within all OATPs are several putative

n-glycosylation sites within the extracellular loop 2 (summarized in Hagenbuch and Meier, 2004). The

latter have been shown to play a role in functionality and membrane targeting of the protein (Lee et

al., 2003). Another conserved feature is the presence of ten cysteine residues in the extracellular loop

5 which seem to be crucial for protein function (Hagenbuch and Meier, 2003; Hänggi et al., 2006).

Moreover the OATP “superfamily-signature” (D-X-RW-(I,V)-GAWW-X-G-(F,L)-L), which is located at

the extracellular border of loop 3 and TMD 6, is highly conserved (summarized in Hagenbuch and

Meier, 2003).

Figure 2: Topological model of human OATP1B1. Based on computational calculations 12 transmembrane domains, five intracellular loops (I1-5) and six extracellular loops (E1-6) were predicted. Conserved cysteine residues are given in gray, N-glycosilation sites are indicated with “Y”. Modified from (Hagenbuch and Meier, 2004).

Transported Substrates

OATPs have a wide and overlapping spectrum of amphipathic endogenous and exogenous substrates

with a high molecular weight of >450 Da. Endogenous substrates are mainly bile salts, hormones and

their conjugates and anionic oligopeptides (summarized in Hagenbuch and Meier, 2003). Examples for

Page 22: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Organic anion transporting polypeptides (OATPs)

- 17 -

bile salts and their conjugates transported by OATPs are taurocholate, cholate or glycocholate, which

are built in the liver from cholesterin. To fulfill their role as emulgator of fatty acids to digest these,

they have to be transported into the bile and then the digestive system. Since 95 % of the bile acids are

recycled by the body and undergo the enterohepatic circulation, they have to reach the portal venous

system and finally the liver again. Since most OATPs transport at least taurocholate, they seem to play

an important role within that bile acid homeostasis along with other transporters. Moreover, OATPs

play an important role in distribution of various hormones within the body. These include steroid

hormones and their conjugates like dehydroepiandrosteron (DHEAS), estrone-3-sulfate (E3S) and

estradiol 17-β-glucuronide (E17βG), which are produced from cholesterol as well as thyroid hormones

like triiodthronin (T3) and thyroxin (T4), which play an important role in the energy metabolism.

Another group of transported hormones are eicosonaids like prostaglandins which contribute as

neurotransmitters to inflammations. Peptides and other transported organic anions are for example

enkephalins, bromosulfophthalein (BSP) or bilirubin. Additionally, OATPs can transport a number of

xenobiotic compounds including pharmaceutical drugs and toxins. Drugs include the antibiotic

rifampicin; pravastatin and fluvastatin, which are used as therapy against high cholesterin values; the

cardiac glycosides quabain and digoxin; cytostatics like methotrexate (MTX), paclitaxel and docetaxel

or the HIV-proteaseinhibitor ritonavir. The cyclopeptid phalloidin from the death cap (Amanita

phalloides) and the cyanotoxin MC are toxins so far found to be transported by OATPs (summarized in

Hagenbuch and Meier, 2003).

In mammals, transport of MC could be demonstrated for the human OATP1B1, OATP1B3 and

OATP1A2 with an affinity of 7±3 µM, 9±3 µM and 20±8 µM respectively, whereas OATP2B1 did not

exhibit a transport for MC (Fischer et al., 2005). In rat, OATP1B2 was found to be responsible for the

uptake of MC into the liver (Fischer et al., 2005). The same transporter was also responsible for MC

uptake in mice, as shown in OATP1B2 null mice, which did not exhibit hepatotoxicity upon MC

exposure compared to control mice expressing the transporter. Nevertheless, MC seems to be taken up

into the liver of these OATP1B2 null mice anyhow as demonstrated in immunoblot analyses (Lu et al.,

2008). Either MC was able to passively pass the membrane to some extend or other transporters are

responsible. To our knowledge, other mammalian OATP transporters have not been investigated

referring their MC transport yet.

Some OATPs, including members of the OATP1A and OATP1B subfamily, have a wide spectrum of the

mentioned substrates, suggesting that they play an overall role in drug absorption and disposition

together with the P-glycoproteins (Mdr) and the multidrug resistance associated proteins (Mrp). Other

transporters however have narrower substrate specificity. Other than the rest of OATP1 family,

OATP1C1 transports BSP,E17βG and E3S to some extend but has a high affinity for the thyroid

hormones T4 and reverse T3 (the biological inactive metabolite of T3). Even though some transporter

Page 23: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Organic anion transporting polypeptides (OATPs)

- 18 -

have the same substrate and one transporter can be responsible for the uptake of various compounds,

affinities can vary quite a bit, suggesting that a transporter in a given tissue has its specific role which

cannot fully be compensated by another transporter (Hagenbuch and Meier, 2004; Hagenbuch and

Stieger, 2013).

OATPs can have multiple binding sites for one or more substrates. OATP1B1 for example transports

E3S with two different affinities (Noé et al., 2007; Tamai et al., 2001), whereas only one binding side

could be found for fluvastatin (Noé et al., 2007). Similarly it has been shown that some transporter like

OATP1A4 transport TCA, digoxin and E17βG. While TCA and digoxin inhibit each other’s transport,

E17βG does not interfere with the transport of the other two substrates, it indeed even stimulates the

uptake of TCA, suggesting that this substrate uses another binding side and is not in competition with

the uptake of TCA or digoxin (Sugiyama et al., 2002).

Tissue distribution

Most OATPs are expressed in multiple tissues like the blood brain barrier, choroid plexus, lung, heart,

liver, intestine, kidney, placenta and testes, but some are solely expressed in specific organs. A

restricted expression in the liver has been found for the human transporter OATP1B1 and OATP1B3 as

well as for the rat OATP1C1 which was specific for the brain, whereas OATP2A1, OATP3A1, OATP4A1

mRNA has been found in all tissues analyzed (Roth et al., 2012; Sugiyama et al., 2002).

In general, OATPs in the liver have been suggested to play a role in the overall uptake and elimination

of endo- and xenobiotics. For example members of the OATP1B subfamily transport a wide range of

substrates but their expression is restricted to the basolateral membrane of hepatocytes, which

suggests that they play an important role in the hepatic clearance of amphipathic compounds. The

human transporter OATP1A2 exhibits the widest range of different substrates. Even though it is

expressed in different organs, the main expression was found in the blood brain barrier, suggesting

that it plays an important role in the uptake of neuroactive peptides and drugs into the brain as well as

in the removal of organic compounds from the brain. Similarly, OATP1C1 mRNA was only found in the

brain and testes. Due to its high affinity of thyroid hormones it has been hypothesized, that it pays a

role in delivery and disposition of thyroid hormones in the mentioned tissues (summarized in

(Hagenbuch and Meier, 2004).

Transport mode and driving force

OATPs are sodium independent, which means unlike other transporter e.g. Na+/taurocholate

cotransporting polypeptide (NTCP) they do not couple a sodium import to transport bile salts. In

general, their transport mode was suggested to function as organic anion exchange, whereby the

substrate is transported in exchange for a counter ion (summarized in (Hagenbuch and Gui, 2008).

Page 24: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Organic anion transporting polypeptides (OATPs)

- 19 -

However, the kind of counter ion seems to depend on the subtype. Mainly bicarbonate and reduced

glutathione and glutathione conjugates have been suggested (Li et al., 1998; Li et al., 2000; Satlin et

al., 1997). Hence, this is the third possible role GSH occupies in the MC toxicity pathway.

Uptake has moreover demonstrated to be pH dependent as shown for several OATPs, which were

stimulated by a low extracellular pH. This stimulation was accompanied by an increased affinity to the

substrates. It was demonstrated that a highly conserved histidine in the third TMD is responsible for

this pH sensitivity, since it is conserved in all OATP subtypes, which showed increased affinity at low

pH but not in the pH-insensitive OATP1C1. An exchange of the glutamine found in this position in

OATP1C1 to histidine resulted in a pH sensitive transport of OATP1C1 as well (Leuthold et al., 2009).

Oatps in fish

The first Oatp in fish was cloned from the cartilaginous species little skate (Leucoraja erinacea) and

classified as Oatp1d1 (Cai et al., 2002; Hagenbuch and Meier, 2004). The 76 kDa protein revealed

highest amino acid identity with the human OATP1C1, however the multispecific transport as well as

the predominant expression at the basolateral membrane of hepatocytes was more similar to

OATP1B1 and OATP1B3. Oatp1d1 is also the only transporter found so far in fish which is transporting

MC. Thereby the affinity is with 26.9±9.9 µM similar to the affinity of the mammalian OATPs (Cai et al.,

2002; Meier-Abt et al., 2007).

Boaru et al. (2006) tested in rainbow trout primary hepatocytes cytotoxicity of MC-LR. A reduced

viability after exposing cells for 48 h to 250 nM MC-LR was nearly reversed when rifampicin was used

for co-incubation. Therefore they concluded that the transport might be Oatp-mediated. They

additionally found an elongated sequence tag (EST) in the rainbow trout database which was 81 %

similar to the Oatp1d1 sequence from little skate. This sequence was used for real time PCR showing

an expression in fresh isolated liver, however not in the primary hepatocytes (Boaru et al., 2006).

In zebrafish, a bioinformatical study contributed to the first identification of several Oatps in zebrafish

(Popovic et al., 2010). Thereby, 14 different Oatps have been found. Later, Oatp1d1 from zebrafish was

fully characterized regarding phylogeny, transported substrates and possible functional sites (Popovic

et al., 2014; Popovic et al., 2013). In accordance with the little skate Oatp1d1, the zebrafish Oatp1d1

was suggested to be a multispecific transporter.

The mRNA expression level of Oatps in fathead minnow (Pimephales promelas) was measured in order

to investigate alterations of Oatp expression upon exposure with thyroid hormones. Therefore,

partially sequenced ESTs similar to other annotated Oatp members were used. However no further

characterization regarding this Oatps has been conducted and sequences are not available with their

whole open reading frame (ORF) (Muzzio et al., 2014).

Page 25: Molecular cloning, characterization and relevance for

GENERAL INTRODUCTION Goal of this study

- 20 -

All other Oatp sequences in fish which can be found in common genome browser including NCBI,

ensemble or zfin are predicted sequences which have been annotated through an automated

computation system, functional analysis however is lacking.

GOAL OF THIS STUDY Aim of this study was the identification and characterization of Oatp subtypes in various fish species,

especially referring their MC transport. As shown above, a lot is known about the dynamics in fish and

the pathological alterations caused by MC. Kinetics however still remain unclear in a lot of areas. Since

several OATPs/Oatps have been demonstrated to transport MC in humans, rat, mice and cartilaginous

fish, they are most likely expressed in bony fish as well and play a key role in MC uptake and

distribution. We propose that bioavailability as well as organ targeting of MC strongly depends on the

expression level and distribution of several Oatp subtypes. This will give us further insight into

toxicokinetics of MC in fish and provides an understanding for the MC mediated toxicity and target

organs of MC. Therefore an Oatp was cloned from rainbow trout and functionally characterized as we

hypothesized that a homologue of the little skate Oatp1d1 and the human OATP1B1/OATP1B3 is

present in the liver of bony fish, which is able to transport MC (CHAPTER 2). Moreover we aimed to

investigate whether species specific differences might be explained by their difference in their

expression level and distribution of these Oatps. Thus, a comparison of the Oatp expression in various

species was conducted (CHAPTER 3). Finally, we wanted to test the hypothesis whether other

transporters than Oatp1d1 are able to transport MC in fish and if variations in uptake properties of

several MC-congeners occur. To answer that question a screening of all annotated zebrafish Oatps was

conducted and functional expressed subtypes were tested for transport of MC-congeners in a time-

and concentration dependent manner (CHAPTER 4).

Page 26: Molecular cloning, characterization and relevance for

CHAPTER 2

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW

TROUT LIVER OATP KONSTANZE STEINER, BRUNO HAGENBUCH AND DANIEL R. DIETRICH

PUBLISHED IN TOXICOLOGY AND APPLIED PHARMACOLOGY

ABSTRACT Cyanobacterial blooms have an impact on the aquatic ecosystem due to the production of toxins (e.g.

microcystins, MCs), which constrain fish health or even cause fish death. However the toxicokinetics of

the most abundant toxin, microcystin-LR (MC-LR), are not yet fully understood. To investigate the

uptake mechanism, the novel Oatp1d1 in rainbow trout (rtOatp1d1) was cloned, identified and

characterized. The cDNA isolated from a clone library consisted of 2772 bp containing a 2115 bp open

reading frame coding for a 705 aa protein with an approximate molecular mass of 80 kDa. This fish

specific transporter belongs to the OATP1 family and has most likely evolved from a common ancestor

of OATP1C1. Real time PCR analysis showed that rtOatp1d1 is predominantly expressed in the liver,

followed by the brain while expression in other organs was not detectable. Transient transfection in

HEK293 cells was used for further characterization. Like its human homologues OATP1A1, OATP1B1

and OATP1B3, rtOatp1d1 displayed multi-specific transport including endogenous and xenobiotic

substrates. Kinetic analyses revealed a Km value of 13.9 µM and 13.4 µM for estrone-3-sulfate and

methotrexate, respectively and a rather low affinity for taurocholate with a Km value of 103 µM.

Furthermore, it was confirmed that rtOatp1d1 is a MC-LR transporter and therefore most likely plays a

key role in the susceptibility of rainbow trout to MC intoxications.

Page 27: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Introduction

- 23 -

INTRODUCTION Cyanobacterial blooms occur worldwide in fresh and coastal water. Since a huge amount of toxins can

be released during or after the breakdown of the bloom, toxic cyanoblooms have been associated with

fish kills all over the globe (Albay et al., 2003; Bürgi and Stadelmann, 2002; Rodger et al., 1994). Fish

kills may be the result of numerous simultaneously occurring factors including e.g. oxygen depletion,

alkaline pH, and cyanotoxins. It is not surprising that cyanobacterial blooms have deleterious effects

on fish populations given that approximately 75 % of the blooms contain toxins (Lawton and Codd,

1991) while at the same time fish, in comparison to terrestrial animals, have limited possibilities in

avoiding exposure to the bloom and its toxins. In fact it could be shown that bony fish from surface

waters experiencing blooms also incorporated the most abundant cyanobacterial toxin present:

microcystin (MC) (Freitas de Magalhães et al., 2001; Krishnamurthy et al., 1986; Sipiä et al., 2001).

Given that there is a fish species-dependent sensitivity towards MC (Kotak et al., 1996; Råbergh et al.,

1991; Tencalla, 1995; Tencalla et al., 1994), these species differences in toxin susceptibility will lead

to profound effects on the species diversity within a given ecosystem. In addition, as discussed

previously in various studies (Peng et al., 2010; Poste et al., 2011), accumulation of MC in edible fish is

likely to pose a threat to human health. However, despite the serious implications described above

only limited information is available e.g. the underlying mechanisms (toxicodynamics) of MC toxicity

in bony fish, while the disposition of MC i.e. MC uptake and systemic distribution in fish remains

completely unclear. Investigation of the latter however could help to understand why certain species

appear more likely to accumulate MC and whether edible parts of the fish are potentially more

contaminated with the toxin than others.

Indeed, experimental applications of either MC containing bloom material or purified MC in fish

resulted in liver-, kidney-, gut- and gill-pathology as well as associated effects e.g. decreased

hemoglobin, increased serum liver enzyme activities and ROS formation (Fischer and Dietrich, 2000;

Li et al., 2004; Tencalla and Dietrich, 1997). These pathological alterations are explained by the fact

that MCs inhibit cytosolic as well as nucleosolic serine (Ser) and threonine (Thr) protein phosphatases

(PP). Ser/Thr-PP-inhibition leads to protein hyperphosphorylation and as a consequence to altered

signal transduction pathways, cytoskeletal disintegration and oxidative stress via mitochondrial

toxicity, finally leading to apoptosis and cell necrosis. MCs are structurally complex molecules with a

molecular weight around 1000 Da. Obviously these amphiphilic to lipophilic cyclic peptides do not

readily pass through the cell membrane to enter into a hydrophilic environment e.g. the cytosol.

Indeed, MCs require active uptake from the cell ambient media and carrier mediated transport across

the membrane into the cell, which was demonstrated in humans, mice, rats and in the little skate

Page 28: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Introduction

- 24 -

(Leucoraja erinacea) to occur via organic anion transporting polypeptides (OATPs/Oatps) (Fischer et

al., 2010; Fischer et al., 2005; Hagenbuch and Gui, 2008; Lu et al., 2008; Meier-Abt et al., 2007).

Organic anion transporting polypeptides (OATPs in humans/ Oatps in fish, summarized in (Hagenbuch

and Stieger, 2013) mediate the uptake of various amphipathic endogenous and exogenous organic

compounds. So far, ≥ 300 OATPs/Oatps have been annotated, thereby representing ≥ 40 animal

species. Based on their amino acid identity OATPs can be divided into six families (OATP1-6, with

≥ 40 % amino acid identity) and several subfamilies (with ≥60 % amino acid identity, summarized

(Hagenbuch and Meier, 2004). OATPs can be expressed in a ubiquitous or in an organ specific manner

(summarized in (Hagenbuch and Stieger, 2013). Typical substrates are bile salts, eicosanoids, steroids

and steroid conjugates, thyroid hormones, prostaglandins, oligo-peptides as well as pharmaceuticals

and toxins, e.g. the phalloidins and amanitins of the highly toxic amanita species (death cap) and the

cyanobacterial microcystins (König, 2011; Roth et al., 2012). In humans MC organ/cellular uptake is

mediated by the liver specific OATP1B1 and OATP1B3 as well as by OATP1A2(Fischer et al., 2010;

Fischer et al., 2005; Monks et al., 2007). In rodents OATP1B2 was identified as a specific MC

transporter (Fischer et al., 2005; Lu et al., 2008).

Based on the knowledge in mammals we hypothesize that an ortholog of the OATP1 subfamily is

expressed in bony fish as well, fulfilling a similar function in the liver and playing a key role in

detoxification. We propose that this ortholog transports MC which would help to understand the

kinetics of MC. We furthermore aim to investigate if the expression profile of this Oatp can explain the

often observed pathologies and accumulations in various fish tissue including the muscle as most

edible tissue.

In cartilaginous fish a transporter belonging to the OATP1 family has been identified and was

demonstrated to transport [3H]-demethylphalloidin and [3H]-dehydro-MC-LR (Cai et al., 2002; Meier-

Abt et al., 2007). In addition, several Oatps have been annotated for zebrafish (Danio rerio) in a

phylogenetic analysis (Popovic et al., 2010). Amongst these, the zebrafish Oatp1d1 was further

characterized regarding its functional and structural properties using specific OATP substrates, albeit

without specifically characterizing its MC transport capabilities (Popovic et al., 2013). As primarily

salmonids appeared to be involved in many toxic cyanobacterial bloom associated fish kills reported,

the widespread salmonid aquaculture, and the reported MC intoxication of salmonids in aquacultures

(net-pen disease, (Andersen et al., 1993), MC contaminated fish may pose an important route of MC

exposure for humans. Therefore the focus for identifying potential Oatp orthologs of the human liver

OATP1 members was placed on salmonids, in this case rainbow trout (Oncorhynchus mykiss). Thus, a

rainbow trout liver cDNA library was screened for a putative Oatp that could be functionally expressed

in vitro and then characterized with regard to Oatp substrate specificity and MC transport.

Page 29: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Methods

- 25 -

METHODS

Reagents and materials

[3H] taurocholic acid (TCA, 0.21 Tera Becquerel (TBq)/mmol), [3H] estrone sulfate ammonium salt

(E3S, 2 TBq/mmol), [3H] methotrexate disodium salt (MTX, 1.8 TBq/mmol), [3H]

estradiol-17-β-D-glucuronide (E17βG, 1.8 TBq/mmol), [3H] bromosulfophthalein [BSP]

(0.5 TBq/mmol), [3H] [D-penicillamine 2,5]encephalin (DPDPE. 1.7 TBq/mmol), [3H]

dehydroepiandrosterone sulfate sodium salt (DHEAS, 2.9 TBq/mmol), [3H] ritonavir (0.04 TBq/mmol),

[3H] paclitaxel (1.7 TBq/mmol), [3H] docetaxel (2.2 TBq/mmol), [3H] pravastatin (0.6 TBq/mmol), [3H]

ouabain (0.6 TBq/mmol), [3H] oleic acid (1.2 TBq/mmol) and [3H] digoxin (1.4 TBq/mmol) were

purchased from Perkin Elmer Inc. (Waltham, USA), American Radiolabeled Chemicals Inc. (Saint Louis,

USA), Moravek Biochemicals (Brea, USA) and Hartmann Analytic GmbH (Braunschweig, Deutschland).

MC-LR was from Enzo Life Science, Inc. (New York, USA). Reverse-transcription and PCR reagents

were purchased from New England Biolabs (Ipswich, UK) and cell culture material from PAA

Laboratories (Cölbe, Germany) unless indicated otherwise. All other chemicals and antibodies, unless

indicated otherwise, were from Sigma-Aldrich (Taufkirchen, Germany). Live rainbow trout were

purchased from a local hatchery (Riebel, Reichenau, Germany).

Isolation of Oatp cDNA and phylogenetic analysis

A rainbow trout liver cDNA library was constructed in the vector pCMV-Sport6 with the Gateway

Super Script Plasmid Kit (Invitrogen, Carlsbad, CA, USA) following the manufacturer’s instructions. The

library was screened with a 32P dCTP labeled PCR-fragment of a rainbow trout sequence, which was

amplified using degenerate primers binding to OATP/Oatp consensus sequences as suggested in (Cai

et al., 2002). The resulting PCR product was sequenced and identified as Oatp via blast analysis, and

gene specific primers were designed. Amplification with these specific primers resulted in a 377 bp

amplicon which was used for screening of the library under high stringency conditions (final two wash

steps: 0.5 × SSC/0.1 % SDS; 65 °C; 20 min). After two rounds of screening a single clone with a 2772 bp

insert was identified. It contained the full-length open reading frame of a rainbow trout liver Oatp

(rtOatp) flanked by 5’- and 3’-untranslated sequences. The DNA sequence was determined for both

strands (Eurofins MWG GmbH, Ebersberg, Germany) and was submitted to NCBI

(http://www.ncbi.nlm.nih.gov/projects/geo) with Acc# KJ831065. After aligning OATP sequences

with ClustalW, phylogenetic analyses were carried out using MEGA 6.06 software (statistical method:

Neighbor-Joining, Jones-Taylor-Thornton model, pairwise deletions). Branch strength was tested using

bootstrap methodology (5000 replicates). Representative OATP amino acid sequences from

vertebrates including fish were used as listed in SUPPLEMENTAL TABLE 1. According to the above

Page 30: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Methods

- 26 -

analyses the cloned rtOatp was classified as rtOatp1d1. The tree was constructed using iTOL (Letunic

and Bork, 2007).

Organ distribution of rtOatp1d1 via PCR and real time PCR

To determine tissue specific expression of rtOatp1d1, rainbow trout were dissected immediately after

cervical dislocation and tissues stored in RNAlater at -80 °C. For isolation of total RNA, tissue samples

were disrupted in TriFast (Peqlab Biotechnologie GmbH, Erlangen, Germany) using a tissue lyser II

(Quiagen, Hilden, Germany). RNA was extracted using a Quiagen Mini Kit according to manufacturer’s

instructions. One microgram of RNA was used to prepare cDNA using 200 U of SuperScript III Reverse

Transcriptase (Invitrogen, Carlsbad, CA, USA), 2.5 µM oligo (dT), 2.5 µM random hexamers, 500 µM

dNTP, 10 mM DTT and 40 U Rnase Inhibitor. The reverse transcription reaction was incubated at 42 °C

for 5 min followed by 45 min at 50 °C. The reaction was stopped by 15 min of incubation at 72 °C. RNA

was degraded by incubation with 5 U RNase H for 20 min at 37 °C. PCR and real-time PCR were carried

out on 5 individual rainbow trout. PCR with subsequent gel-electrophoresis was carried out with Taq

polymerase and the specific primers for Oatp1d1 as listed in SUPPLEMENTAL TABLE 2. Elongation factor

1α (ef1α) was used as reference gene. For semi-quantitative real time PCR a 1:5 dilution of the cDNA

was used whereby two reference genes, ef1α and 18S-rRNA, served as normalization controls as a

similar expression amongst organs could be shown for the two reference genes (SUPPLEMENTAL FIGURE

1F). SYBR Green (Bioline, London, UK) was used according to manufacturer’s instructions. RNA

melting curve analysis was conducted following real-time PCR to ensure lack of secondary products

(SUPPLEMENTAL FIGURE 1A-C). Standard curves for the primers used were analyzed in a preliminary

study with technical triplicates, whereby efficiency of the amplification was 76 % for 18s-rRNA and

nearly 100 % for ef1α and rtOatp1d1 (SUPPLEMENTAL FIGURE 1D-E). Data were analyzed according to

the Δct method.

Expression of rtOatp1d1 in HEK293 cells

Human embryonic kidney (HEK293) cells were cultured in Dulbecco's Modified Eagle Medium

(DMEM) with 25 mM glucose and 4 mM L-glutamine supplemented with 10 % FBS, 100 units/ml

penicillin and 100 mg/ml streptomycin at 37 °C and 5 % CO2. For transient transfections of rtOatp1d1

two different plasmids were used. One plasmid was the original pCMV-Sport6 containing the isolated

cDNA. The other plasmid was constructed by cutting the cDNA with NotI and EcoRI out of pCMV-

Sport6 and cloning the resulting insert after gel purification into the precut pTracerTMCMV/Bsd

(Invitrogen, Carlsbad, CA, USA). This construct allowed using expression of GFP to verify transfection

success. Comparison of transfection with both vectors with regard to taurocholate uptake resulted in

similar transport results (data not shown). The respective cDNA or the empty vector as a control

Page 31: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Methods

- 27 -

(0.5 µg per well on a 24-well plate) were transfected with FuGENE (Promega) at 60-80 % HEK293 cell

confluence.

Uptake studies

HEK293 cells were seeded on poly-D-lysine coated 24 well plates and transfected as described above.

Substrate uptake was measured 24 h and 48 h after transfection with either pCMV-Sport6 or

pTracerTMCMV/Bsd, respectively. For uptake experiments cells were washed 3 times with warm

uptake buffer (116.4 mM NaCl, 5.3 mM KCl, 1.6 mM NaH2PO4, 0.8 mM MgSO4, 5.5 mM D-Glucose,

20 mM Hepes, pH adjusted to 7.4 with Trizma base). Subsequently, cells were incubated with

200 µl/well of uptake buffer containing 0.3 µCi of the respective radiolabeled substrate at the

following final concentrations: [3H] TCA (30.0 nM), [3H] E3S (6.6 nM), [3H] MTX (13.9 nM),

[3H] E17βG (6.1 nM), [3H] BSP (20.7 nM), [3H] DPDPE (6.7 nM), [3H] DHEAS (3.8 nM),

[3H] ritonavir (300.0 nM), [3H] paclitaxel (6.6 nM), [3H] docetaxel (5.0 nM) or

[3H] pravastatin (40.0 nM). After 5 min of incubation substrate uptake was stopped by washing the

cells 6 times with ice-cold uptake buffer. Subsequently, cells were solubilized with 300 µl 1 % Triton X-

100 per well and 200 µl was used to quantify radioactivity by β-scintillation counting, while the

remaining solution was used to determine total protein concentrations (triplicate analyses) using the

BCA protein assay kit (Pierce Biotechnology, Rockford, USA).

For three demonstrated rtOatp1d1 substrates additional experiments were performed to determine

uptake kinetics under initial linear rate conditions (30 s) at substrate concentrations ranging from

0.1 µM to maximally 300 µM. Uptake was normalized to total protein concentration within each

respective experiment. Transporter specific uptake was calculated by subtracting the counts of empty

vector transfected cells from those of rtOatp1d1-transfected cells. Data points were means ± SEM of

three technical replicates and a minimum of 4 experiments were performed. In one case, only 1

experiment could be performed. In that case the 3 technical replicates served to determine the means

± SEM for each time-point. A non-linear regression (Michaelis-Menten) curve was fitted to the data to

obtain the Km-values.

Immunoblotting for MC-LR detection

Forty-eight hours after transfection with pTracerTMCMV/Bsd HEK293 cells were incubated with 50 nM

MC-LR for 1 h, 6 h or 24 h. Incubation with methanol served as solvent control, whereby methanol

concentrations never exceeded 0.2 %. HEK293 cells stably expressing human OATP1B3 served as

positive and empty vector transfected cells or non-transfected cells as negative controls. Proteins were

isolated from cells with cold buffer containing 10 mM Tris-base, pH 7.5, 140 mM NaCl, 5 mM EDTA,

0.1 % (v/v) Triton X-100, pH 7.5 and 1 % protease inhibitor. Following centrifugation at 10,000 g,

Page 32: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Methods

- 28 -

protein content in the supernatant was measured using the BCA assay kit. Fifteen µg of protein/lane

was separated using 10 % sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) at

constant 200 V (Laemmli, 1970). Proteins were then transferred onto a nitrocellulose membrane

(Whatman, Dassel, Germany) at 300 mA for 90 min (Towbin et al., 1979) and membranes were

incubated in blocking buffer (Tris-buffered saline with1 % Tween 20 containing 1 % bovine serum

albumin) for 1 h at room temperature. To detect MC-LR bound to phosphatases, nitrocellulose

membranes were incubated with the MC-specific primary monoclonal anti-Adda- antibody (1:10) over

night at 4 °C (clone AD4G2, (Zeck et al., 2001)). To detect the primary antibody the membranes were

incubated with horseradish peroxidase (HRP)-conjugated goat anti-mouse antibody (1:80,000; Sigma-

Aldrich, Taufkirchen, Germany) for 1 h at room temperature. Immunopositive bands were visualized

with Amersham ECL AdvanceTM Western Blotting Detection Kit (GE Healthcare, Amersham, UK)

according to manufacturer’s protocol and the chemiluminescent signal was detected using

ImageQuant LAS 4000 mini (GE Healthcare, Germany). After stripping the membranes with 0.2 M

NaOH for 10 min they were re-probed with the polyclonal rabbit anti-GAPDH antibody (1:200, Santa

Cruz Biotechnologies, Heidelberg, Germany) again using overnight 4 °C incubation. GAPDH served as

immunoblot as well as protein reference control. Quantification was carried out with grayscale

analysis (Quantity One version 4.6.9).

Immunoblotting for rtOatp1d1

A polyclonal rabbit antibody was generated against a peptide corresponding to 13 amino acids

(PRKSDCDRMFKFYM) at the C-terminal end of rtOatp1d1 by Washington Biotechnology, Inc.,

Baltimore, USA.

Rainbow trout liver tissue (0.5 g) was homogenized in a glass-teflon potter with 10 ml homogenization

buffer (0.25 M Sucrose, 10 mM Tris-HCl, 1 % protease inhibitor and 40 µg/ml

phenylmethylsulfonylfluorid). The homogenate was centrifuged for 5 min at 1,000 X g and 4 °C. The

resulting supernatant was centrifuged for 5 min at 6,000 x g and 4 °C. The pellet was discarded and the

supernatant was centrifuged for 2 h at 100,000 x g and 4 °C. The resulting pellet containing the

membrane protein fraction, was re-suspended in 500 µl buffer (0.25 M sucrose and 20 mM HEPES at a

pH 7.5) and 25 µg protein were separated using a 10 % gel. SDS-PAGE and western blotting were

carried out as described above. Nitrocellulose membranes were blocked with 2 % milk powder in Tris-

buffered saline with 1 % Tween 20 overnight at 4 °C followed by incubation with either a 1:10,000

dilution of the anti-rtOatp1d1 antiserum or pre-immune serum for 2 h at 4 °C. Detection and

visualization was carried out after 1 h of incubation with a horseradish peroxidase (HRP)-conjugated

goat anti-rabbit antibody (1:160,000, Sigma- Aldrich, Taufkirchen, Germany) and subsequent

visualization was obtained with the Amersham ECL AdvanceTM Western Blotting Detection Kit (GE

Page 33: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Methods

- 29 -

Healthcare, Amersham, UK). The chemiluminescent signal was detected using ImageQuant LAS 4000

mini (GE Healthcare, Germany).

Page 34: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Results

- 30 -

RESULTS

Cloning and phylogenetic analysis of a cDNA encoding rtOatp1d1

To identify and characterize a rainbow trout ortholog of the human, mouse, rat and little skate

OATPs/Oatps, which have been demonstrated to transport MC, a 32P-labeled 377 bp DNA fragment

was used as a probe to screen a cDNA library obtained from rainbow trout liver. A 2772 bp cDNA was

isolated containing a 5’-UTR of 91 bp, an open reading frame (ORF) of 2115 bp and a 3’-UTR of 586 bp.

The ORF encodes a protein with 705 amino acids and based on phylogenetic analyses clusters within

the OATP1 family (FIGURE 3). The rainbow trout Oatp amino acid sequence shares 53-60 % identity

with the annotated fish-specific Oatp1d1 subfamily but only 45-48 % identity with members of the

OATP1C1 subfamily also present in fish and other vertebrates. Based on these analyses this new

rainbow trout Oatp was classified as Oatp1d1, the sequence was deposited (Acc# KJ831065) and is

subsequently called rtOatp1d1.

Figure 3: Phylogenetic tree of selected vertebrates in the OATP1 family. Based on a ClustalW alignment the tree was assembled using Neighbor-Joining Method with pairwise deletion; bootstrap values (5000 replicates) > 0.5 are indicated as gray circles, while the size is proportional to the bootstrap value. * indicates predicted sequences. Fish Oatps are marked in gray, the novel rtOatp1d1 is highlighted with a light gray background.

Tissue distribution of rtOatp1d1

Tissue distribution of rtOatp1d1 at the RNA level was determined using semi-quantitative qRT-PCR.

Data summarized in (FIGURE 4A) revealed a predominant expression of rtOatp1d1 in the liver, while

Page 35: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Results

- 31 -

its expression was moderate in the brain. No mRNA was detected in kidney, muscle and intestine,

either due to absence of rtOatp1d1 or due to too low a level of expression to allow reliable detection.

Western blot analyses were used to confirm that rtOatp1d1 indeed is expressed in the liver (FIGURE

4B). An immunopositive band at 80 kDa was detected corresponding nicely with size predictions

based on the amino acid composition of rtOatp1d1 sequence (77.1 kDa). Low expression levels

prevented detection of rtOatp1d1 protein in other tissues.

Figure 4: Organ distribution of rtOatp1d1. (A) RNA expression of rtOatp1d1 in several rainbow trout tissues performed with semi-quantitative real time PCR. Expression of rtOatp1d1 was normalized to the mean expression of elongation factor 1α and 18s-RNA as reference genes. Whisker box plots with minimum to the maximum value whiskers; n = 5 fish. (B) rtOatp1d1 protein expression in rainbow trout liver homogenate. Lane 1 pre-immune serum; Lane 2: anti-rtOatp1d1 antibody.

rtOatp1d1 substrate specificity

To test the substrate specificity of rtOatp1d1, transiently transfected HEK293 cells (rtOatp1d1-

HEK293) were used to determine the uptake of 11 radiolabeled compounds i.e. reported substrates for

human OATP1B1, OATP1B3 and OATP1A2. FIGURE 5 demonstrates that rtOatp1d1-HEK293 cells

transported pravastatin, E3S, E17βG, TCA, BSP, DPDPE, DHEAS and MTX, whereby uptake rates ranged

between 0.02 (Pravastatin) and 0.44 (TCA) pmol/mg/protein/min. No transport was observed for

docetaxel, paclitaxel and ritonavir.

Page 36: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Results

- 32 -

Figure 5: Functional characterization of rtOatp1d1 in HEK293 cells. 24 h after transfection, the cells were incubated for 5 min with [3H] TCA (30.0 nM), [3H] E3S (6.6 nM), [3H] MTX (13.9 nM), [3H] E17βG (6.1 nM), [3H] BSP (20.7 nM), [3H] DPDPE (6.7 nM), [3H] DHEAS (3.8 nM), [3H] ritonavir (300.0 nM), [3H] paclitaxel (6.6 nM), [3H] docetaxel (5.0 nM) or [3H] pravastatin (40.0 nM). Values represent the mean ± SEM of technical triplicates. Differences in substrate uptake between rtOatp1d1-HEK293 (gray bars) and either eV-HEK293 (in case of DPDPE, DHEAS, TCA and MTX) or non-transfected cells (all others, represented in white bars). Significant differences were statistically analyzed using a non-parametric Mann–Whitney test (* p<0.05).

E3S, TCA and MTX kinetics in rtOatp1d1-HEK293 cells

Kinetic experiments were carried out to further characterize substrate transport mediated by

rtOatp1d1. Concentration dependent uptake was performed for E3S (0.1 – 75 µM), MTX (0.1 – 75 µM)

and TCA (0.1 - 300 µM) under initial linear rate conditions (30 s) (FIGURE 6). Nonlinear regression

analyses (Michaelis-Menten) revealed Km values of 13.9 µM, 13.4 µM and 103 µM for E3S, MTX and

TCA, respectively.

Page 37: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Results

- 33 -

Figure 6: Kinetic analyses of rtOatp1d1-mediated substrate uptake. Transport was measured within the initial linear uptake phase at 30 s with varying concentrations between 0.1 µM and 75 µM or 300 µM respectively. (A) E3S (B) TCA and (C) MTX. Values depict mean ± SEM of five (A) or four (B) independent replicates with three technical replicates each. For MTX (C) three determinations of one replicate are shown. A nonlinear regression (Michaelis-Menten) was fitted to the net uptake which was calculated by subtracting the unspecific uptake of the eV-HEK293 cells from the uptake of the rtOatp1d1-HEK293 cells.

Page 38: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Results

- 34 -

Microcystin-LR transport by rtOatp1d1-HEK293 cells

In order to test whether rtOatp1d1 is capable of transporting MC-LR, rtOatp1d1-HEK293 cells,

HEK293 cells transfected with the empty vector as a negative control, and HEK293 cells expressing

human OATP1B3 as a positive control were incubated with 50 nM MC-LR for 1, 6 or 24 h. Subsequent

western-blotting with the AD4G2 MC-antibody demonstrated MC covalently bound to intracellular

phosphatases and thus represented the MC-LR transported by rtOatp1d1 or OATP1B3 into the

HEK293 cells. Indeed, a strong MC-immuno-positive band was detected (between 30 and 40 kDa) for

HEK293 cells expressing either rtOatp1d1 or human OATP1B3 (FIGURE 7A) which compared well with

similar immunoblots of OATP1B3 transfected HEK293 cells (Fischer et al., 2010) and neuronal cells

(Feurstein et al., 2009) exposed to MC-LR. As previously demonstrated by Feurstein et al. (2009) the

MC-LR immunopositive bands corresponded to MC-LR covalently bound to the catalytic subunit of PP1

and PP2A, albeit no specific immuno-detection of PP1 and 2A was carried out on the same blots.

Quantification of the immunoblots suggested that immunodetectable MC-LR increased with MC-LR

exposure time in both, the transiently transfected rtOatp1d1-HEK293 cells (FIGURE 7B) and the stably

transfected human OATP1B3-HEK293 cells (FIGURE 7C). Detectable MC-LR decreased when

rtOatp1d1-HEK293 cells where co-incubated with 100 µM TCA or BSP (FIGURE 8). Contrary to

expectations, co-incubation with 10 µM TCA or BSP did not lead to a reduced signal, suggesting a

higher affinity of MC-LR than BSP or TCA for rtOatp1d1, as also suggested by kinetic experiments with

TCA. As demonstrated earlier by Fischer et al. (2010) TCA did not appear to significantly compete with

MC-LR uptake in OATP1B3-HEK293 cells. Co-incubation with BSP however provided for a similar

reduction of detectable intracellular MC-LR in both rtOatp1d1- and OATP1B3-transfected HEK293

cells.

Page 39: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Results

- 35 -

Figure 7: MC-LR uptake mediated by rtOatp1d1. (A) Immunoblot showing uptake of MC-LR mediated by transiently transfected rtOatp1d1-HEK293, ev-HEK293, non-transfected cells (non-tr.) or stable transfected human OATP1B3-HEK293. The cells were incubated with 50 nM MC-LR (+), or solvent control MeOH (-) for 1, 6 and 24 h, GAPDH has been used as reference gene; (B and C). Densitometric grayscale analysis (n=5, mean ± SEM) was used to compare the uptake of MC-LR in rtOatp1d1 (B) and human OATP1B3- transfected HEK293 cells (C), after subtracting grayscale signals for ev-HEK293 cells or non-transfected HEK293 cells respectively.

Figure 8: Competitive inhibition of rtOatp1d1-mediated uptake of MC-LR by TCA and BSP. Immunoblot showing uptake of 50nM MC-LR for 24 h (MC-LR +) in the absence of presence of 100 µM or 10 µM TCA or BSP. MeOH was used as solvent control (-). rtOatp1d1: transiently transfected rtOatp1d1-HEK293 cells; OATP1B3: stably transfected OATP1B3-HEK cells.

Page 40: Molecular cloning, characterization and relevance for

DISCUSSION In this study a novel rainbow trout Oatp transporter (rtOatp) was identified and its functional

characterization revealed that it can transport MC. Indeed, screening of a rainbow trout liver cDNA

library resulted in the isolation of a 2772 bp cDNA that encodes for a 705 aa protein. Subsequent blast

analysis confirmed the presumed affiliation of this sequence with the OATP1 family. The ensuing

phylogenetic analysis of OATP1 subtypes from 16 vertebrates including 9 fish species added more

insight and suggested that the novel rtOatp belongs to the Oatp1d1 subfamily, which to date has only

been reported in fish.

Functional expression of rtOatp1d1 in HEK293 cells resulted in the transport of several of the known

OATP substrates, including E17βG, E3S, TCA, DPDPE, DHEAS, MTX, pravastatin, docetaxel, paclitaxel

and MC (summarized in(Hagenbuch and Stieger, 2013), suggesting that rtOatp1d1 is a multi-specific

transporter comparable with the liver-specific human OATP1B1 and OATP1B3 or the human

OATP1A2, primarily expressed in the brain but also in various other tissues. rtOatp1d1 affinities for

the three tested substrates E3S, TCA and MTX were in general comparable to the affinities of other

transporter of the OATP1 family in humans and fish, which mostly vary between 10 and 80 µM (TABLE

1). However, given that these results were obtained by different research groups in vastly differing

expression systems, the comparison of the substrate affinities is not absolute, but can serve for

orientation purposes. It appears that within fish, both little skate and rainbow trout Oatp1d1 have

comparable affinities for TCA, while they vary for E3S (TABLE 1). Transport of E3S in rtOatp1d1

appears to be monophasic with a Km value of 13.9 µM, whereas a biphasic transport mechanism has

been shown for OATP1B1 (Gui and Hagenbuch, 2009; Noé et al., 2007; Tamai et al., 2001).

Several members of the OATP1 subfamily have been demonstrated to transport various MC congeners

(Fischer et al., 2010). Based on the fact that this novel rtOatp1d1 is also a member of the OATP1

subfamily it is not surprising that also rtOatp1d1 is capable of transporting MC-LR. A similar MC-LR

uptake velocity was found when comparing the time course of MC-LR uptake into HEK293 cells

transiently transfected with rtOatp1d1 and stably transfected with human OATP1B3 (FIGURE 7B and

C). Direct competition experiments using BSP and TCA to inhibit MC-LR uptake (FIGURE 8) suggest that

MC-LR has a lower affinity for rtOatp1d1 than for OATP1B3 given that the affinities of TCA for the two

transporter are similar (TABLE 1). In view of the lacking Km values for BSP in rtOatp1d, the affinity of

BSP and thus competition with MC-LR can only be speculated, albeit the comparison of western blots

(FIGURE 8) would suggest that BSP has a lower affinity to rtOatp1d1 when tested at equimolar

competitive concentrations of MC-LR. The fact that other Oatp substrates like TCA and BSP can inhibit

MC uptake conversely raises the question whether MC is also able to minimize the uptake of

physiological substrates transported by Oatp1d1 with a low affinity. Since this could lead to an altered

homeostasis of endogenous substrates, a reduced fitness of the fish would be expected.

Page 41: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Discussion

- 37 -

Table 1: Affinities for the tested substrates in this study compared to human and fish transporter of the Oatp1 family. Given as Km value in µM.subtype species Km [µM] references

E3S TCA MTX

OATP1B1 Human 0.23/45 10-34 - *1

OATP1B3 Human 58 6-42 25-39 *2

OATP1A2 Human 16-59 60 457 *3

Oatp1d1 Zebrafish 1.75 - - *4

Oatp1d1 Little skate 61 85 - *5

Oatp1d1 Rainbow trout 13.9 103 13.4

*1 (Abe et al., 1999; Cui et al., 2001; Hsiang et al., 1999; Noé et al., 2007),

*2 (Abe et al., 2001; Briz et al., 2006; Gui et al., 2008; Leuthold et al., 2009),

*3 (Badagnani et al., 2006; Bossuyt et al., 1996; Kullak-Ublick et al., 1995; Lee et

al., 2005),*4 (Popovic et al., 2013), *5 (Cai et al., 2002).

In order to understand the significance of rtOatp1d1 for potential MC induced organ toxicity,

expression analyses suggested that rtOatp1d1 is highly expressed in the liver followed by the brain.

This distribution pattern compares well with the ortholog pattern found in the little skate Oatp1d1

(Cai et al., 2002). Thus rtOatp1d1 represents a functional fish homologue of the mammalian

OATP1B1/OATP1B3 with a similar substrate spectrum i.e. including the cyanotoxin MC. Therefore a

functional role of rtOatp1d1 as transporter for bile acid metabolism, steroid distribution and

detoxification can be assumed.

The organ distribution pattern observed and the capability to transport MC-LR corroborate earlier

findings of intra-organ/intra-cellular MC detected in Salmonidae and Cyprinidae (Fischer and Dietrich,

2000; Fischer et al., 2000). Pathologies of MC were mainly reported in the liver of several fish,

suggesting that due to high expression of Oatp1d1 in the liver, the liver is primarily affected during MC

intoxications. However, MC was also reported in the brain of carp (Cyprinus carpio, (Fischer and

Dietrich, 2000) and live-bearing fish (Cazenave et al., 2005), suggesting an Oatp1d1 mediated MC

accumulation and neurotoxicity. Indeed, changes in fish swimming activity (Cazenave et al., 2008;

Ernst et al., 2007; Tencalla et al., 1994) support a MC mediated neurotoxicity. Consequently the

presence of fish Oatp1d1 in liver and brain appears to be a key for MC uptake and the subsequent

development of MC mediated organ pathology and thus morbidity and mortality of fish exposed.

Interestingly organs primarily affected following MC intoxication as well as the severity of

pathologies/toxicity observed varies between species. Indeed, while in rainbow trout pathology was

Page 42: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Discussion

- 38 -

primarily observed in the liver, cyprinids presented also with kidney pathology, albeit less severe than

that observed in the liver of the same animals (Carbis et al., 1996; Fischer and Dietrich, 2000;

Råbergh et al., 1991). The latter observations corroborate findings of a species-specific susceptibility

towards MC upon oral exposure and intraperitoneal injection and employing LD50/LC50 values for

comparison (Fischer and Dietrich, 2000; Kotak et al., 1996; Råbergh et al., 1991; Tencalla, 1995;

Tencalla et al., 1994). As the expression levels and MC transporting capabilities of OATPs/Oatps

appear more critical than the ser/thr PP inhibiting capabilities of the individual MC-congeners for the

development of tissue damage and thus the apical toxicity (Fischer et al., 2010), current data would

suggest that also for fish the expression and expression level of Oatp1d1 is the critical factor governing

morbidity and mortality of fish exposed to MC containing cyanobacterial blooms. Hence, adverse

effects of toxic cyanobacterial blooms to aquatic ecosystems including a shift of species distribution

largely depend on the organ distribution of Oatp1d1 and their inherent susceptibility to MCs of the

respective inhabitant fish species.

It is interesting to note that oral exposure of fish to MC resulted in pathological alterations and

accumulation of MC in various organs including kidney, muscle gut and gill (Fischer and Dietrich,

2000; Li et al., 2004; Tencalla and Dietrich, 1997). Despite the obvious transport of MCs across the

gastro-intestinal wall to the portal vein, expression of rtOatp1d1 was not found in the gastro-intestinal

tract of rainbow trout. A very low expression level of rtOatp1d1 and thus low uptake of MC per time

unit would be in disagreement with the rapid onset of pathology in the liver as observed upon oral MC

exposure in trout and carp. Likewise an accumulation of MC in muscle tissue cannot be explained by

the rtOatp1d1 organ expression data shown. As most likely other Oatp subtypes are present in the fish,

the presence of MC in the gastro-intestinal tract and muscle can be explained by a so far unidentified

transporter, capable of transporting MCs. Indeed, several studies reported the accumulation of MC in

fish muscle tissue at tissue concentrations that would surpass the tolerable daily intake (TDI) value,

recommended by the world health organization (WHO) (Freitas de Magalhães et al., 2001; Poste et al.,

2011; Song et al., 2007), upon normal daily fish consumption.

In summary, despite that rtOatp1d1 shows the highest phylogenetic relationship with OATP1C1, the

comparable functional characteristics places it closer to the human OATP1A2, OATP1B1 and

OATP1B3. Indeed, the broad spectrum of transported substrates, including bile salts, hormone

steroids and xenobiotica, overlaps greatly with the human OATP1A and OATP1B subtypes suggesting

that rtOatp1d1 plays a similar role in fish as OATP1A2, OATP1B1 and OATP1B3 in humans. Similar to

these three human OATPs rtOatp1d1 was identified as an MC transporter, allowing for further insight

into the kinetics of MC in fish. Based on the organ distribution of rtOatp1d1 in rainbow trout, adverse

effects mediated by acute as well as chronic MC exposure may be expected primarily in the liver and in

Page 43: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND FUNCTIONAL CHARACTERIZATION OF A RAINBOW TROUT LIVER OATP

Funding

- 39 -

brain and result in the acute or slow demise of individual fish as well as whole populations in

ecosystems subject to recurring toxin producing cyanobacterial blooms.

FUNDING This work was enabled by financial support from a DFG Center Grant, the International Max Planck

Research School for Organismal Biology (IMPRS), by a Marie Curie International Research Staff

Exchange Scheme Fellowship (PIRSES-GA-2011-295223), and by grants from the National Center for

Research Resources (RR021940) and the National Institute of General Medical Sciences (GM077336

and GM103549) of the National Institutes of Health.

Page 44: Molecular cloning, characterization and relevance for

CHAPTER 3

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN

WHITEFISH AND COMMON CARP KONSTANZE STEINER, HEINKE BASTEK, BRUNO HAGENBUCH AND DANIEL R. DIETRICH

IN PREPARATION

ABSTRACT Microcystins (MC) are hepatotoxins produced by multiple cyanobacterial species worldwide. During

cyanobacterial blooms high concentrations of this heptapeptide occurs in the water. This poses not

only a threat to terrestrial animals, but also to aquatic animals, especially fish which only have limited

possibilities to avoid the bloom and toxin. Death and severe pathologies after chronic exposure to MCs

have been demonstrated in several fish species, whereby severity seems to differ depending on the

species. Uptake of MC into the target cells is mainly mediated by several subtypes of organic anion

transporting polypeptides (OATPs), as shown in humans and rodents. In fish, this type of transport has

only been shown for the bony fish rainbow trout (Onchorhynchus mykiss) and the cartilaginous fish

little skate (Leucoraja erinacea). In both species the Oatp subtype identified belonged to the fish

specific Oatp1d1 subfamily. Knowledge about Oatps in other fish, especially non-model fish species is

limited. The aim of this study was to use molecular cloning in concert with sequencing to identify the

transporter Oatp1d1 from the two species common carp (Cyprinus carpio) and whitefish (Coregonus

wartmanni), both non-model species and present in the lake of Constance. Novel sequences where

identified and contained an open reading frame of 2061 base pairs (bp) for common carp and 2116 bp

for whitefish. Classification of the two transporter sequences within the fish specific Oatp1d1

subfamily was confirmed by phylogenetic and alignment analysis. Using real time PCR the highest RNA

expression of Oatp1d1 was identified in the liver of both species. In whitefish the only other organ

which is expressing Oatp1d1 was the brain, whereas in the common carp the kidney also contained

Oatp1d1 RNA. These species specific expression pattern suggests differences in MC organ toxicology

and highlight the challenges in understanding or predicting the impact of this natural toxin on aquatic

ecosystems.

Page 45: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Introduction

- 43 -

INTRODUCTION Cyanobacteria can be found all over the globe in fresh and coastal water. Favorable conditions

including nutrient enrichment and increased temperature are likely to trigger growth leading to mass

occurrence (also called blooms) in water bodies including recreational, drinking and agricultural

water resources (Manganelli et al., 2012). These blooms are accompanied with an unpleasant odor,

water colorations and declines in water quality including oxygen depletion, which can cause the death

of various animals including fish (Jewel et al., 2003). More than 50 cyanobacterial species produce

toxic metabolites of various chemical structures (Sivonen and Jones, 1999) and these toxic metabolites

have a huge impact on terrestrial and aquatic animals (Martins and Vasconcelos, 2009; Stewart et al.,

2008). Acute exposure to cyanotoxins can lead to fish kills. Sub-acute and chronic exposure can greatly

impact the fresh water ecosystems because it can lead to a shift in species distribution due to varying

sensitivity of the aquatic species to the sub lethal toxin doses.

The most abundant and persistent group of cyanotoxins are the cyclic heptapeptide microcystins (MC),

of which more than 100 congeners have been described. It is therefore not surprising, that

accumulation of MC in fish tissues has been demonstrated in numerous studies. Interestingly, a few of

these studies found variations in the MC-content in different species obtained from the same lake

(Amrani et al., 2014; Chen et al., 2009; Ernst et al., 2001; Schmidt et al., 2013; Singh and Asthana,

2014). This might be explained by manifold ecological niches (diets, habitats) as well as by anatomical

alterations, especially of the digestive system. The latter is of particular importance because the main

uptake route for fish is oral ingestion of toxin contaminated food with subsequent digestion via the

intestine (Bury et al., 1995; Tencalla et al., 1994). Target organs, particularly the liver, are then

reached via the blood stream, leading to decreased haemoglobin, increased serum liver enzyme

activities and reactive oxygen species (ROS) formation (Fischer and Dietrich, 2000; Li et al., 2004;

Tencalla and Dietrich, 1997). Even though MC is mainly described as hepatotoxin, tissue damage has

also been described for the kidney and gills (Ernst et al., 2007; Fischer and Dietrich, 2000; Råbergh et

al., 1991).

However, several laboratory studies showed differences in the median lethal dose (LD50) of MC after

oral and intraperitoneal (ip) injection (Kotak et al., 1996; Råbergh et al., 1991). Since fish have been

exposed to similar concentrations of MC, these differences in LD50 cannot be explained by the above

factors. It is therefore more likely that alterations in enzymatic and protein expression are responsible

for variations of the bioavailability of MC as well as for MC uptake into the target organs. Due to its

structure and size, MC can only enter the cells via a specific transport system. It has been shown that

organic anion transporting polypeptides (humans OATP, fish Oatp) are responsible for the MC uptake

Page 46: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Introduction

- 44 -

in humans, rodents and fish (CHAPTER 2, Fischer et al., 2005; Meier-Abt et al., 2007). Therefore, the

expression of Oatps likely plays a critical role in the distribution of MC.

OATPs/Oatps are a group of membrane solute carriers, which have been identified in more than 40

different species so far (summarized in Hagenbuch and Stieger, 2013). Based on their amino acid

sequence identities, OATPs/Oatps have been classified into several families and sub -families.

OATPs/Oatps with more than 40 % identity belong to the same family (e.g. OATP1, OATP2), while

Oatps with more than 60 % identity belong to the same subfamily (e.g. OATP1A, OATP1B, OATP1C,

summarized in (Hagenbuch and Meier, 2004). While some OATPs/Oatps mediate uptake of only a few

substrates, others have a wide substrate spectrum, which can overlap between the different

transporter. OATPs/Oatps can be expressed in an ubiquitous or an organ-specific way (summarized in

Hagenbuch and Stieger, 2013).

When MC enters the cell via this carrier-mediated transport, toxicity is induced through binding of MC

to protein phosphatase (PP) 1 and 2a. Since these enzymes play a key role in the phosphorylation

balance in the cell, this inhibition results in altered signal transduction pathways, cytoskeletal

disintegration, oxidative stress and finally apoptosis and cell necrosis (summarized in(Campos and

Vasconcelos, 2010).

The bioavailability and accumulation in the cell of MC and strongly depends on the toxicokinetics and

therefore the amount of MC transported into the cell. Thus, the Oatp mediated uptake is a limiting

factor. We hypothesize that the expression level and organ distribution of MC transporting

OATPs/Oatps are critical to species sensitivity because higher OATP/Oatp expression levels probably

result in a higher uptake rate of MC. Previously it has been shown that Oatp1d1 is responsible for MC

uptake in rainbow trout (Oncorhynchus mykiss)and little skate (Meier-Abt et al., 2007). This subtype is

a fish specific transporter, which so far has been found and further characterized regarding

transported substrates in the bony fish zebrafish (Danio rerio), rainbow trout and the cartilaginous

fish, little skate (Cai et al., 2002; Popovic et al., 2013). In the model organism zebrafish several other

Oatp subtypes besides Oatp1d1 have been identified, however, no screening of MC transport or other

further analysis was conducted (Popovic et al., 2010). A few additional Oatp1d1 sequences as well as

other subtypes from other fish species were annotated based on automated computational analysis,

without further functional analysis. However, it is not known, if these transporters are also expressed

in other non-model fish species. Common carp (Cyprinius carpio) and whitefish (Coregonus

wartmanni) are two species of interest since they are often found to be affected by MC exposure. By

comparing their Oatp organ distribution we also consider representatives of the two major families

Salmonids and Cyprinids. In this study we aim to identify the full length Oatp1d1 sequence in these

species in order to analyze the phylogenetic relationship of the Oatps across the different fish species.

Page 47: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Introduction

- 45 -

Furthermore, we compared the organ distribution, since differences in these might explain the species

specific sensitivity.

Page 48: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Material and Methods

- 46 -

MATERIAL AND METHODS

Fish and reagents

Common carp were purchased at a local hatchery (Riegger, Ettenheim) and whitefish were caught

from Lake Constance by a professional fisherman. Enzymes and chemicals for molecular biology

applications were from New England Biolabs (Ipswich, UK) unless otherwise indicated, all other

chemicals were from Sigma-Aldrich (Taufkirchen, Germany).

Total RNA isolation and cDNA preparation

Organs were immediately stored in RNA later (Quiagen, Hilden, Germany). Upon dissection of fish,

organs were immediately stored in RNA later (Quiagen, Hilden, Germany). To isolate total RNA,

samples were homogenized in peqGOLD TriFastTM (Peqlab Biotechnologie GmbH, Erlangen, Germany)

using a tissue lyser (Quiagen, Hilden, Germany). Then 1/5 volume of chloroform was added and the

samples were centrifuged at 8000 x g and 4 °C for 15 min. The chloroform phase was mixed with an

equal volume of 70 % ethanol and then added to Quiagen columns (Quiagen, Hilden, Germany).

Washing and elution was carried out as described in the manufacturer’s manual. The RNA

concentration and purity were determined spectrophotometrically. The integrity of the RNA used for

the organ distribution was assessed using formaldehyde agarose gel electrophoresis. RNA was

converted into cDNA using 200 U reverse transcriptase super script® III (Invitrogen, Carlsbad, CA,

USA), 2.5 µM oligo (dT), 2.5 µM random hexamer, 500 µM dNTP, 10 mM DTT and 40 U Rnase Inhibitor.

Reverse transcription was run at 42 °C for 5 min followed by 50 °C for 45 min and 72 °C for 15 min.

RNA was degraded by incubation with 5 U RNase H for 20 min at 37 °C.

Isolation of carp and whitefish Oatp cDNA

For the isolation of a carp Oatp cDNA, mRNA was isolated using the PolyATtract mRNA isolation

system (Promega, Fitchburg, WI, USA) and used to construct a liver cDNA library with the Gateway

Super Script Plasmid Kit following the manufacturer’s instructions. To screen this cDNA library a

400 bp DNA probe was amplified from carp liver cDNA with gene specific primers designed based on

the information available from another cyprinid, the zebrafish Oatp1d1. After sequencing and

confirming that the cDNA corresponded to an OATP sequence it was labeled using the DIG high prime

DNA labeling and Detection Starter Kit II (Roche, Basel, Schweiz) and used for screening according to

manufacturer’s instruction. High stringency conditions (hybridization at 42 °C overnight; post-wash

steps: 2 × saline-sodium-citrate (SSC)/0.1 % sodium dodecyl sulfate (SDS) at 20 °C for 2 x 5 min and

0.5 x SSC/0.1 % SDS at 68 °C for 2 x 15 min) were used to ensure identifying specific products. The

whitefish Oatp was isolated using the rapid amplification of cDNA ends (RACE)-PCR technique. For

Page 49: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Material and Methods

- 47 -

that, cDNA was transcribed from total liver RNA as described above. With this method a first, unknown

part of the sequence is getting identified with degenerate primers. Those were designed based on an

alignment of known Oatp1d1 sequences (KJ831065, NP_001082802 and AAL66021). By sequencing

the amplification product, gene specific primers can be designed in the known part of the sequence,

while degenerate primers help to identify the sequence further in 5´- and 3´-direction. As untranslated

regions (UTR) of a specific sequence in general are not similar enough between species, the RACE

System for rapid amplification of cDNA was used according to manufacturer’s instruction for

identifying the 5´-start and the 3´-end (3´ RACE- and 5´ RACE System for Rapid Amplification of cDNA

Ends by Life Technologies, Carlsbad, CA, USA). DNA sequences were verified for both strands (Eurofins

MWG GmbH, Ebersberg, Germany) and revealed a 2528 kb insert for the carp Oatp and a 2653 kb

insert for the whitefish Oatp. Both inserts contained the respective full length ORF.

Phylogenetic analysis

Amino acid sequence alignments were generated with representative protein sequences from

vertebrates including fish (SUPPLEMENTAL TABLE 3) with the ClustalW algorithm using ClustalW2

(http://www.ebi.ac.uk/Tools/msa/clustalw2/) and following multiple alignment options: Protein

weight matrix: Gonnet; gap open: 10; gap extension: 0.2. Sequences in SUPPLEMENTAL TABLE 3 and in

the phylogenetic trees are named according to their published annotation with the following

exceptions: (1) The green spotted puffer fish Oatp (CAF89838) can be found in the NCBI database as

unnamed protein.As it clusters within other Oatp1d1 from fish (data not shown), we annotated it as

Oatp1d1. However, this sequence was only used for the alignment in SUPPLEMENTAL TABLE 3 but not

for phylogenetic analysis presented in this study, since the first 17 amino acids are missing and we

aimed to only use full length open reading frames (ORF). (2) The protein sequence XP_416418 from

chicken was annotated as “OATP1C1-like”, as it clusters with OATP1B3 sequences we re-annotated it

as OATP1B3. Predicted sequences, which were annotated using computational methods, are marked

with a star and were tested for introns using GENSCAN (http://genes.mit.edu/GENSCAN.html).

Phylogenetic analysis was performed in MEGA 6.06 with the Neighbor-Joining method as statistical

method (Substitution model: Dayhoff model, pairwise deletions). To test the phylogeny, bootstrap

methodology (1000 replicates) was applied. According to these analyses both novel sequences were

classified as Oatp1d1. Trees were drawn using iTOL (Letunic and Bork, 2007). Transmembrane

domains were predicted using TOPCONS (http://topcons.cbr.su.se/), Potential N- linked glycosylation

sites were predicted by NetNGlyc 1.0 (http://www.cbs.dtu.dk/services/NetNGlyc/) and

phosphorylation sites for PKC and PKA were predicted by GPS 2.0 (Xue et al., 2008).

Page 50: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Material and Methods

- 48 -

Organ distribution of carp and whitefish Oatp1d1 using semi-quantitative real time PCR

Semi-quantitative real time PCR was conducted using a 1:5 cDNA dilution as template and SensiMix™

SYBR® & Fluorescein Kit (Bioline, London, UK) according to manufacturer’s instructions. For carp,

RNA from five fish was analyzed and for whitefish RNA from nine individuals was analyzed. In order to

find a suitable reference gene, which is equally expressed in all organs, we tested EF1α, 18S ribosomal

RNA and β-actin for all fish and chose Ef1α for whitefish and 18S ribosomal RNA for carp as reference

gene (data not shown). Building of primer-dimers and unspecific products could be excluded with a

melt curve analysis subsequent to the PCR. Data were analyzed according to the ΔΔct method.

Standard curves for the primers used (SUPPLEMENTAL TABLE 4) were analyzed for each biological

replicate, whereby efficiency of the amplification was 106 ± 17 % for whitefish EF1α, 92 ± 4 % for

whitefish Oatp1d1, 113 ± 9 % for carp 18S ribosomal RNA 134 ± 28 % and for carp Oatp1d1 (data not

shown).

Page 51: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Results

- 49 -

RESULTS

Cloning and phylogenetic analysis of cDNA encoding Oatp1d1

In order to identify possible candidate transporters which might be responsible for MC transport into

fish hepatocytes we screened a carp liver cDNA library and used RACE PCR for whitefish. One Oatp

cDNA sequence was identified in both species. In carp it consisted of 2528 basepairs (bp) with an open

reading frame of 2061 bp encoding a 687 amino acid protein. From whitefish liver RNA a 2653 bp PCR

product was amplified. The open reading frame was 2116 bp long and encoded a 705 amino acid

protein. In order to classify the novel sequences to an Oatp subfamily and give an overview of the

phylogenetic relationships of OATPs/Oatps in mammals and fish, a phylogenetic tree was generated.

This tree contained representatives of all OATP families from mammals as well as all so far annotated

and predicted ORFs of Oatps in fish. Sequences encoding for OATPs/Oatps of 12 fish species belonging

to various families of bony and cartilaginous fish were found in the genome browser Ensembl and

NCBI. As shown in FIGURE 9, these fish Oatps cluster within all known human OATP families, with the

exception of the OATP6 family where no fish ortholog has been found so far. The novel sequences

cloned in this study belong to the Oatp1d subfamily, which is restricted to fish species and cannot be

found in mammals. In order to better characterize the phylogenetic relationship in the OATP1/Oatp1

family to which Oatp1d1 belongs, we calculated a tree using amino acid sequences of OATPs/Oatps in

family 1. The subfamilies Oatp1f, Oatp1d, OATP1A, OATP1B and OATP1C/Oatp1c built distinct clusters

which are supported by the robustness of the tree as bootstrap values at the referring branch were

> 98 % (with the exception of Oatp1c1 from african coelacanth and Oatp1d1 from little skate which is

discussed later). OATP1C1/Oatp1c1 is present in both mammals and fish, whereas the OATP1B and

OATP1A subfamilies are restricted to mammals. In contrast, the Oatp1f and Oatp1d subfamilies can

only be found in fish. As expected, the novel Oatp1d1 sequence cloned from whitefish clustered with

Oatp1d1 cloned from rainbow trout, as both fish belong to the salmonid family. Likewise the two

cyprinid Oatp1d1 sequences from carp and zebrafish clustered together (FIGURE 10). The novel

Oatp1d1 sequences share common Oatp features as shown in the alignment in FIGURE 11, including

the Oatp-signature at the border of extracellular domain 3 and transmembrane domain 6 and the

Kazal-type serine protease inhibitor domain in the big extracellular loop 5. Eight of the ten conserved

cysteine residues are found within the Kazal domain. Hydrophobicity analyses predicts 12

transmembrane domains and consensus sequences for 4 glycosylation sites and 7 PKC/PKA

phosphorylation sites can be found in at least 6 of the 8 aligned sequences. Notably, when calculating

distances based on the alignments, all Oatp1d1 sequences show an amino acid identity of at least

60 %, which was the defined threshold for subfamilies (Hagenbuch and Meier, 2004), only the little

skate Oatp1d1 shares less than 50 % amino acid identity with all other sequences (TABLE 2)

Page 52: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF

OATP1D1 IN WHITEFISH AND COMMON CARP

Results

- 48 -

Figure 9: Phylogenetic analysis of representative organic anion transporting polypeptides (OATPs/Oatps) from mammals and fish. The tree is based on a ClustalW alignment and was built using Dayhoff model method with pairwise deletion; bootstrap values (1000 replicates) > 60 % are indicated as gray circles, while the size of the circle is proportional to the bootstrap value. The tree was calculated using Mega6 software. * indicates predicted sequences based on automated computational analysis. Oatps in fish are marked in blue, the novel sequences are highlighted with a yellow background. The numbers represent the classification into the six OATP-families, while the letters indicate the subfamilies.

Page 53: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Results

- 51 -

Figure 10: Phylogenetic analysis of representative mammalian and fish Oatps belonging to the family 1based on a ClustalW alignment. The tree was assembled using Dayhoff model method with pairwise deletion; bootstrap values (1000 replicates) > 60 % are indicated. The line segment represents the length of the branch with 0.1 substitutions per site. * indicates predicted sequences. Oatps in fish are marked in blue, the novel sequences are highlighted with a yellow background.

Page 54: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Results

- 52 -

Figure 11: Amino acid alignment of full length Oatp1d1 sequences in several fish species based on ClustalW alignment. Members of the Oatp1d1 family, including the two novel Oatp1d1 members, contain 12 transmembrane domains as indicated with numbers. The Kazal domain and the Oatp-signature, two common Oatp features are indicated with the black boxes. Conserved cysteine residues are shown in bold. N-glycosylation sites are marked with stars, phosphorylation sites for PKC and PKA which are conserved in at least six sequences are marked with triangles.

Page 55: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Results

- 53 -

Table 2: Amino acid identities [%] of the Alignment shown in Figure 11, based on ClustalW-algorithm.

Whi

tefis

h

Rain

bow

trou

t

Zebr

afis

h

Com

mon

car

p

Gree

n sp

otte

d pu

ffer

fish

Japa

nese

puf

fer

fish

Stic

kleb

ack

Litt

le s

kate

Whitefish 100 93.3 68.0 68.9 68.2 68.6 64.5 49.8

Rainbow trout 93.3 100 68.0 68.7 67.3 67.6 64.2 49.6

Zebrafish 68 68.0 100 84.2 66.7 66.2 63.8 49.4

Carp 68.9 68.7 84.2 100 67.4 68.3 64.2 49.0

Green spotted puffer fish 68.2 67.3 66.7 67.4 100 87.4 73.4 50.1

Japanese puffer fish 68.6 67.6 66.2 68.3 87.4 100 72.1 50.2

Stickleback 64.5 64.2 63.8 64.2 73.4 72.1 100 48.5

Little skate 49.8 49.6 49.4 49.0 50.1 50.2 48.5 100

Tissue distribution of Oatp1d1 in various fish species

Based on the fact, that Oatp1d1 of rainbow trout and little skate can transport MC (chapter 2),

(Meier-Abt et al., 2007), and since MC toxicity has been documented in various fish organs (Ernst et al.,

2006; Fischer and Dietrich, 2000; Tencalla and Dietrich, 1997), we determined an expression profile

of the Oatp1d1-mRNA in whitefish and carp and tried to compare the expression levels to rainbow

trout and zebrafish (original data for rainbow trout in chapter 1, for zebrafish in chapter 3). Semi-

quantitative real time PCR was used to compare the expression of Oatp1d1 relative to a reference gene

which was equally expressed in all tested organs of one fish species. In whitefish, rainbow trout, and

carp Oatp1d1 is predominantly expressed in the liver (FIGURE 12A-C). To better demonstrate low

expression levels, the ordinate is scaled in two segments. This shows an additional minor expression

in the brain in the mentioned fish species. As Oatp1d1 expression in the brain of carp and rainbow

trout was below 1 % of the house keeping gene, we additionally performed PCR for second verification

(SUPPLEMENTAL FIGURE 2). In carp, expression was additionally observed in the kidney. In zebrafish,

Oatp1d1 seems to be expressed equally in the liver and brain. A quantitative comparison of expression

levels among species was not possible since housekeeping genes were not expressed to equal levels

(SUPPLEMENTAL FIGURE 3A-D).

Page 56: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Results

- 54 -

Figure 12: Expression-levels of Oatp1d1-mRNA in various organs of common carp (A), whitefish (B), rainbow trout (C) and zebrafish (D) were determined using semi-quantitative real-time PCR. Expression of Oatp1d1 relative to the housekeeping gene used in the respective fish species. Data were analyzed using the Δct method. For common carp (n=5) and zebrafish (n=3) 18s-rRNA has been used as reference gene, for whitefish (n=9) and rainbow trout (n=5) EF1α has been used as reference gene. All experiments were conducted in technical triplicates.

Page 57: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Discussion

- 55 -

DISCUSSION The transport of MC is mediated by members of the OATP gene superfamily as shown in humans,

rodents and fish (Fischer et al., 2005; Lu et al., 2008; Meier-Abt et al., 2007). These transporters

therefore play an important role in the distribution of the toxin within organisms. In fish, however,

only limited information on existing Oatp subtypes is available. In order to test if similar Oatps exist in

other fish species than rainbow trout and little skate as well, we cloned and sequenced Oatp1d1 from

the common carp and the whitefish, non-model fish species which can be found in the Lake of

Constance (Germany). Furthermore, we compared the organ distribution of these Oatps with their

distribution in rainbow trout and zebrafish. Using the amino acid sequences of the two newly

identified Oatps together with the already published fish Oatp sequences, we provide an overview of

annotated and predicted Oatp sequences in various fish species, and discuss their place in an

evolutionary context.

Oatps present in fish and phylogenetic analysis

All full length amino acid sequences in fish annotated as Oatps as of August 2014 are shown in the

phylogenetic tree in FIGURE 9. Most of these sequences were predicted by automated computational

analysis; therefore further information including organ distribution and functional analysis about

these Oatps are not available. Amino acid sequences of the model organism zebrafish have been

identified by comparing human OATP sequences with the genome of zebrafish (Popovic et al., 2010).

Popovic et al. (2010) proposed that 14 Oatp subtypes exist, however, only 11 could be found in NCBI.

In the little skate as well as in the rainbow trout an Oatp1d1 was identified by molecular cloning (Cai

et al., 2002) just like Oatp1d1 from carp and whitefish in this study. The tree demonstrates that all

OATP families which are found in humans are also represented in fish, suggesting that an ortholog of

all those families have already been present before the fish/mammalian split. The family OATP6 builds

an exception as no member of this family could be found in fish so far.

As expected, the novel Oatp1d1 sequences identified in this study, show high identity with Oatp1d1

sequences from other fish species and clearly cluster within this subfamily. Oatp1d1 from whitefish

and rainbow trout, two fish belonging to the salmonids, share a very high amino acid identity of

93.3 %. Likewise, the two Oatp1d1 sequences from carp and zebrafish, both cyprinids, cluster next to

each other with an amino acid identity of 74 %. All other Oatp1d1 sequences exhibit an identity with

each other of at least > 60 %, indicating a high conservation of this transporter sequence throughout

bony fish species. An exception constitutes the ortholog of the little skate, which exhibits less than

50 % amino acid sequence identity to all other sequences. This Oatp in little skate has been annotated

as Oatp1d1 by Hagenbuch and Meier (2004) and re-annotated as Oatp1c1 by Popovic et al. (2013). Our

Page 58: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Discussion

- 56 -

analysis with a bootstrap support of < 50 % demonstrates that it cannot clearly be classified within

either the Oatp1c1 or the Oatp1d1 subfamily. The low bootstrap support could be explained by the fact

that little skate is the only fish in the tree which belongs to a different class of fish. While all other fish

belong to the bony fish, little skate is a cartilaginous fish, which evolved earlier in time. As suggested

recently, the threshold for subfamilies (60 %) should not be used as absolute numbers. An example for

that is Oatp1a2 from Xenopus laevis, which is an ortholog of the human OATP1A2, even though it

shares only 48 % amino acid sequence identity (Hagenbuch and Stieger, 2013). Since Oatp1d1 in the

little skate is expressed predominantly in the liver and exhibits multispecific transport, just like other

Oatp1d1 from bony fish, it is tempting to speculate that this functional orthology is based on a

phylogenetic orthology. However, this conclusion cannot generally be made, moreover our

phylogenetic analysis does not clearly support this hypothesis. Thus it should be reconsidered, if the

little skate Oatp1d1 actually represents a member of a novel subfamily.

Within the Oatp1 family, the subfamilies OATP1B and OATP1A are restricted to mammals while

Oatp1d and Oatp1f are restricted to fish. OATP1C1/Oatp1c1 is present in both lineages. Thus, we

assume that OATP1C1 in mammals derived from Oatp1c1 in fish. Whether other transporters, which

are specific for the fish or the mammalian lineage, derived before or after the fish/mammalian split is

not definite. Moreover, low bootstrap values in the referring branches do not clearly support the

location of the subfamily cluster within the tree. Thus, several homologies between the transporters

are conceivable. Referring to the phylogenetic relation between members of the human OATP1 family,

it was suggested that OATP1A1, OATP1B1 and OATP1B3 derived from OATP1C1 through gene

duplication, as OATP1C1 is encoded on the shortest gene (Pizzagalli et al., 2002). Therefore these

transporters would be paralogs to the transporter they derived from. Similar could hold true for fish

Oatp transporters. However, even if no phylogenetic orthology can be proven from the phylogenetic

analysis, a functional similarity between fish Oatp1d1 and members of the mammalian OATP1B

subfamily seems to be present. Both are predominantly expressed in the liver and exhibit multispecific

transport, thus it is likely that they accomplish the same function in the liver.

Species specific organ distribution

Chronic and acute exposure of MC leads to pathological changes and ultimately to the death of fish

species, as shown in field and laboratory studies (Malbrouck and Kestemont, 2006). Investigations,

which were sampling various species and measuring the MC concentration in several tissues suggest

different sensitivity towards the toxin (Chen et al., 2009; Schmidt et al., 2013; Singh and Asthana,

2014). Furthermore, different median lethal dose (LD50) as well as varying organ pathologies have

been found for carp and rainbow trout (Fischer and Dietrich, 2000; Kotak et al., 1996; Råbergh et al.,

1991; Tencalla et al., 1994). Because Oatp1d1 plays a key role in MC transport, we hypothesized that

Page 59: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Discussion

- 57 -

the expression pattern of Oatp1d1 varies between fish species and therefore leads to variations in

uptake properties and thus different MC concentrations in target organs.

We suggest that Oatp1d1 is the main transporter for MC in fish and its expression explains partly the

organ toxicities observed in several studies (Carbis et al., 1996; Ernst et al., 2006; Fischer and

Dietrich, 2000; Råbergh et al., 1991; Tencalla and Dietrich, 1997). In this study we demonstrated that

the whitefish Oatp1d1 has a similar expression pattern to the closely related Oatp1d1 from rainbow

trout. Both mRNAs showed predominant expression in the liver and minor expression in the brain.

The cyprinid common carp additionally had significant levels of Oatp1d1 mRNA in the kidneys,

quantitatively between the mRNA level of Oatp1d1 in liver and in brain. This expression profile is in

accordance with the pathologies demonstrated in carp liver and kidney after ip or oral exposure

(Carbis et al., 1996; Fischer and Dietrich, 2000; Råbergh et al., 1991), whereas in rainbow trout

pathologies were restricted to the liver (Tencalla and Dietrich, 1997). Similarly in whitefish, organ

toxicity was mainly observed in the liver, even though differences in the appearance were also noticed

in the kidneys, but those were not significant and only randomly observed (Ernst et al., 2006). Even

though pathologies in the brain in general could not be seen or were not investigated, accumulation of

MC in the brain has been demonstrated in several field and laboratory studies in the common carp

(Fischer and Dietrich, 2000; Papadimitriou et al., 2012). A change in swimming behavior, which is

regularly observed, provides further evidence for a neurotoxic effect of MC in fish.

A few studies have demonstrated that several other tissues including intestine and muscle can also

contain MC. This has mainly been shown for the common carp, where samples have been taken from

several lakes. Highest amounts of MC were found in liver, followed by the kidney and brain, which

supports our hypothesis that the Oatp1d1 expression level is crucial for the MC amount in the tissue.

Smaller amounts of MC in the intestine and muscle however cannot be explained by an Oatp1d1

uptake since this transporter did not seem to be expressed in these tissues, at least at the mRNA level

(Amrani et al., 2014; Moutou et al., 2012; Papadimitriou et al., 2013). Nevertheless, other studies

observed not only an accumulation of MC but even an accompanied pathological alteration in the

gastrointestinal tract after orally exposing carp to a suspension of freeze-dried Microcystis aeruginosa

(Fischer and Dietrich, 2000). We therefore propose that other Oatp subtypes might play a role in MC

uptake. Indeed, in humans three transporters have been identified as MC transporter. OATP1B1 and

OATP1B3, which are liver specific and OATP1A2 which is more ubiquitously expressed, including the

blood brain barrier. To understand how MC is taken uptake into the intestine or into muscle tissue

further studies are needed and might result in the identification of additional Oatps that can transport

MC.

Page 60: Molecular cloning, characterization and relevance for

MOLECULAR CLONING AND ORGAN DISTRIBUTION OF OATP1D1 IN WHITEFISH AND COMMON CARP

Acknowledgements

- 58 -

In summary, we have identified two novel Oatp1d1 sequences from common carp and whitefish which

might be responsible for MC transport like their orthologs in other fish species. Their organ

distribution is species dependent but is in line with described pathologies and MC accumulation in

carp and whitefish. Although further investigations are needed in order to prove that the Oatp

expression level leads to variations in MC sensitivity, we provide some initial evidence that Oatp1d1

might be involved. Further characterization with respect to accumulation, transport and impact of MC

in additional fish species will help predicting the impact of cyanobacterial blooms on natural

ecosystems and aquacultures.

ACKNOWLEDGEMENTS We would like to thank Steffen Hörer for experimental assistance. This work was enabled by financial

support from a DFG Center Grant and the International Max Planck Research School for Organismal

Biology (IMPRS).

Page 61: Molecular cloning, characterization and relevance for

CHAPTER 4

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-

CONGENERS KONSTANZE STEINER, LISA ZIMMERMANN, BRUNO HAGENBUCH AND DANIEL R. DIETRICH

PUBLISHED IN ARCHIVES OF TOXICOLOGY

ABSTRACT Background: Microcystins (MC), representing >100 congeners being produced by cyanobacteria, are

a hazard for aquatic species. As MC congeners vary in their toxicity, the congener composition of a

bloom primarily dictates the severity of adverse effects and appears primarily to be governed by

toxicokinetics, i.e. whether transport of MCs occurs via organic anion transporting polypeptides

(Oatps). Differences in observed MC toxicity in various fish species suggest differential expression of

Oatp subtypes leading to varying tissue distribution of the very same MC congener within different

species. Objectives: Functional characterization and analysis of the tissue distribution of Oatp

subtypes in zebrafish (Danio rerio) as a surrogate model for cyprinid fish. Methods: Zebrafish Oatps

(zfOatps) were cloned and the organ distribution was determined at the mRNA level. zfOatps were

transiently expressed in HEK293 cells for functional characterization using the Oatp substrates

estrone-3-sulfate, taurocholate and methotrexate and specific MC-congeners (MC-LR, -RR, -LF and -

LW). Results: Novel zfOatp isoforms were isolated. Among these isoforms, organ specific expression of

zfOatp1d1 and of members of the zfOatp1f subfamily were identified. At the functional level,

zfOatp1d1, zfOatp1f2, zfOatp1f3 and zfOatp1f4 transported at least one of the Oatp substrates, and

zfOatp1d1, zfOatp1f2 and zfOatp1f4 were shown to transport MC congeners. MC-LF and MC-LW were

generally transported faster than MC-LR and MC-RR. Conclusions: The subtype specific expression of

zfOatp1d1 and of members of the zfOatp1f subfamily as well as differences in the transport of MC

congeners could explain the MC congener dependent differences in toxicity in cyprinids.

Page 62: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Introduction

- 61 -

INTRODUCTION Mass occurrences of cyanobacteria present a threat to livestock worldwide, since various toxins are

released upon breakdown of the bloom. The composition of the cyanobacteria species and the

produced toxins play an important role in the severity of toxicity (Dietrich et al., 2008; Sivonen and

Jones, 1999). Due to mass-occurrences in nearly all water bodies, cyanobacterial blooms have been

associated with massive kills of fish and other aquatic organisms. Beyond the acute intoxication,

subchronic or chronic exposure to cyanobacterial toxins e.g. microcystins (MCs) may lead to the loss of

the species with the highest susceptibility to cyanotoxins and thus to a shift in the species distribution

of a given aquatic ecosystem (Ernst et al., 2009). MCs are toxic cyclic heptapeptides of which >100

different MC congeners have been described (Puddick et al., 2014). Intoxications by MCs largely

depend on the individual MC congener’s toxicodynamics and toxicokinetics, whereby the latter

depends on the physicochemical properties of the MC congener in question. The toxicodynamic

portion of the toxicity is primarily governed by the capacity of MCs to inhibit Ser/Thr protein

phosphatases (ser/thrPP), consequently resulting in inhibition of crucial cell signal-transduction

processes, protein trafficking, and cytoskeletal homeostasis (Campos and Vasconcelos, 2010). It is

noteworthy that the more lipophilic MCs, namely MC-LW and MC-LF and the more hydrophilic MC-LR

and MC-RR inhibit ser/thrPP activities at approximately equimolar concentrations and thus

physicochemical properties do not explain the large differences observed for the apical toxicity

(Fischer et al., 2010; Höger et al., 2007). Consequently, the key differences in toxicity most likely stem

from differences in toxicokinetics. Indeed, MCs were demonstrated to be transported across cell

membranes via organic anion transporting polypeptides (humans and rodents: OATPs; fish: Oatps

according to HUGO and ZFIN guidelines) (Fischer et al., 2005; Lu et al., 2008; Meier-Abt et al., 2007;

Steiner et al., 2014). To date, more than 300 OATPs/Oatps have been described and different

OATP/Oatp subtypes are found within a single species (Hagenbuch and Stieger, 2013). Computational

analysis and experimental evidence suggest a membrane topology with 12 transmembrane domains

(TMD) for the rat rOATP1A1 (Wang et al., 2008), a feature also common to fish Oatps and other

mammalian OATPs. Although of great importance for the hepatic transport of bile salts, steroids and

steroid conjugates, thyroid hormones, prostaglandins and oligo-peptides as well as for the clearance of

xenobiotics, OATPs are present in almost every organ or tissue. Depending on the subtype, OATPs are

expressed ubiquitously or in specific organs (Hagenbuch and Stieger, 2013).

Transport of MCs was demonstrated for hOATP1A2 which is mainly expressed in the brain and for the

liver specific hOATP1B1 and hOATP1B3 in humans (Fischer et al., 2010; Fischer et al., 2005; Monks et

al., 2007) as well as for rOATP1B2 in rats and mOATP1B2 in mice (Fischer et al., 2005; Lu et al., 2008).

Likewise in fish, MC transport is mediated by liver specific Oatps as shown for Oatp1d1 in little skate

Page 63: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Introduction

- 62 -

and rtOatp1d1 in rainbow trout (Meier-Abt et al., 2007; Steiner et al., 2014). However, MC transport

by OATPs/Oatps is largely MC congener dependent, as demonstrated for hOATP1B1 and hOATP1B3

expressed in HEK293 or HeLa cells with MC-LR, MC-RR, MC-LW, and MC–LF (Fischer et al., 2010;

Monks et al., 2007). Indeed, while some OATPs lacked transport of MC-RR, transport was

demonstrated for MC-LF, MC-LW and to a comparably minor extent for MC-LR (Fischer et al., 2010).

The latter may be key for the systemic distribution of MC congeners within a given organism (fish or

mammalian) and thus to the organ where MC mediated toxicity may evolve. Thus, the differences in

susceptibility toward MC intoxications observed in fish may be explained by differential expression of

Oatp subtypes and by Oatp subtype-selective transport of specific MC congeners. To test the latter

hypothesis, zebrafish (Danio rerio) Oatp subtypes were cloned and expressed in HEK293 cells to allow

characterization of Oatp substrate and MC congener dependent transport.

Page 64: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Material and Methods

- 63 -

MATERIAL AND METHODS

Experimental fish, reagents and materials

Zebrafish (Danio rerio) were obtained from the animal research facility of the University of Konstanz.

[3H] taurocholic acid (TCA) (209.8 Giga Becquerel (GBq)/mmol), [3H] estrone sulfate ammonium salt

(E3S) (2009.1 GBq/mmol) and [3H] methotrexate disodium salt (MTX) (1824.1 GBq/mmol) were

purchased from American Radiolabeled Chemicals Inc. and Hartmann Analytic GmbH (Braunschweig,

Deutschland). MC congeners (-LR, -LF, -LW and -RR) were purchased from Enzo Life Science, Inc. (New

York, USA). Reverse-transcription and PCR reagents were purchased from New England Biolabs

(Ipswich, UK) unless indicated otherwise. Cell culture material was from PAA Laboratories (Cölbe,

Germany). All other chemicals and antibodies, unless stated otherwise, were from Sigma-Aldrich

(Taufkirchen, Germany).

RNA preparation and cDNA Synthesis

RNA was isolated from muscle, intestine, liver, kidney, brain and gills (pooled from 15 zebrafish,

indiscriminant of gender) and cDNA was prepared as described previously in detail (Steiner et al.,

2014).

Cloning of Oatp sequences in eukaryotic expression plasmid

To generate zfOatp expression plasmids, the full-length open reading frames (ORFs) of the Oatp

encoding slco genes were cloned into the vector ptracer™CMV/Bsd (Invitrogen, Carlsbad, CA, USA).

Based on the annotated zfOatp sequences (Popovic et al., 2010), specific primers including flanking

restriction sites were designed to PCR amplify the ORFs of the different zfOatps (SUPPLEMENTAL TABLE

5). cDNA from zebrafish (wild type strain) was used as template and PCR was carried out using

Phusion High Fidelity DNA-Polymerase (NEB) with proofreading activity to exclude that sequence

variations occurred due to methodical reasons. PCR was performed as follows: Initial DNA denaturing

for 30 sec at 98 °C, 35 cycles of denaturing for 10 sec at 98 °C, primer annealing for 30 sec at 52 °C and

elongation for 2.5 min at 72 °C followed by a final extension of DNA-fragments for 5 min at 72 °C.

Before ligation of the vector backbone ptracerTMCMV/Bsd with the ORF, both were digested with the

same restriction enzymes. Transformation of chemical competent E. coli was performed and ligation

success was validated by E. coli colony PCR screening and sequencing.

Page 65: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Material and Methods

- 64 -

Phylogenetic analysis and topology prediction

The resulting cDNAs were sequenced on both strands and aligned with their respective annotated

sequences retrieved from the NCBI database (http://www.ncbi.nlm.nih.gov/). Global alignments of the

two sequences were performed with the ClustalW algorithm. Membrane topology was predicted by

comparing 3 different programs and associated models: TMHMM

(http://www.cbs.dtu.dk/services/TMHMM-2.0/), TMpred (http://embnet.vital-

it.ch/software/TMPRED_form.html) and TOPCONS (http://topcons.cbr.su.se/). The accession

numbers of the zfOatp subtype sequences cloned in this study are listed in SUPPLEMENTAL TABLE 6.

Cell culture and transient transfection

Human embryonic kidney cells (HEK293) were cultured as described by (Steiner et al., 2014). For

expression of zfOatps, 60-80 % confluent HEK293 cells were transiently transfected either with the

empty vector as control (ev-HEK293) or with one of the vector constructs containing the respective

zfOatp ORF using FuGENE transfection reagent (Promega, Madison, WI, USA). The presence of a green

fluorescence protein (GFP) ORF on ptracer™CMV/Bsd allowed verification of transfection success.

Uptake experiments

HEK293 cells were seeded in poly-D-lysine coated 24 well plates and transfected as described above.

48 h post-transfection HEK293 cells were incubated with Oatp substrates for 5 min as previously

described (Steiner et al., 2014). Uptake was normalized to total protein concentration. Experiments

were repeated three times (true replicates) in triplicates each (technical replicates). An Oatp was

considered to transport a substrate if uptake into cells expressing the Oatp was significantly higher

than uptake into control cells which were transfected with the empty vector using the Mann-Whitney

test (GraphPad Prism Software Inc.) with p ≤ 0.05 (one-tailed) from three independent experiments

(n=3). In addition, uptake of [3H] E3S was also performed for up to 30 min with zfOatp1f1 and

Oatp1f2-1 to rule out that [3H] E3S transport could not have been seen due to high affinity/low

capacity transport properties of zfOatp1f1 and Oatp1f2-1. However, no [3H] E3S transport was

observed for these two Oatps (SUPPLEMENTAL FIGURE 4).

Microcystin transport detected via immunoblot

All MC uptake experiments were conducted 48 h post-transfection of HEK293 cells with

ptracerTMCMV/Bsd-Oatp. Time- and concentration dependent uptake of MC-congeners was

determined via immunoblotting. MC congeners were dissolved in methanol (MeOH ≤ 0.2 %) and 0.2%

MeOH served as solvent control. Empty vector transfected HEK293 (ev-HEK293) served as negative

control. For immunoblotting, proteins were isolated from cells with cold buffer containing 10 mM Tris-

Page 66: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Material and Methods

- 65 -

base, pH 7.5, 140 mM NaCl, 5 mM EDTA, 0.1 % (v/v) Triton X-100, pH 7.5 and 1 % protease inhibitor

followed by centrifugation at 10,000 x g. Supernatant protein concentration was measured using a

BCA assay kit (see above). Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) was

performed with 15 µg protein/lane on 10 % gels at 200 V according to Laemmli (1970). Subsequent to

electrophoresis, proteins were transferred onto a nitrocellulose membrane (Whatman, Dassel,

Germany) at 300 mA for 90 min as described elsewhere (Towbin et al., 1979). Membranes were

blocked in Tris-buffered Saline with 1 % Tween 20 containing 1 % bovine serum albumin. To detect

MC bound to phosphatases, nitrocellulose membranes were incubated with the MC-specific primary

monoclonal anti-Adda- antibody (clone AD4G2, 1:10) (Zeck et al., 2001) over night at 4 °C. GAPDH or

actin served as immunoblot as well as protein loading control. Quantification was carried out using

greyscale analysis (Quantity One version 4.6.9).

All cloned zfOatps were screened for their MC transport capability via incubation of the zfOatp-

HEK293 cells with 50 nM of a single MC congener for 24 h. The time-dependent uptake of individual

MC congeners was tested via incubation of the zfOatp-HEK293 cells for 1 h, 6 h and 24 h (and 48 h in

the case of zfOatp1f2) with 50 nM MC (data not shown). Based on the latter results, optimized time-

windows for individual zfOatp and MC congener combinations were selected that allowed time

dependent detection of MC congener transport. For improved kinetic understanding zfOatp1d1-

HEK293 cells were exposed to MC-LR for 3 h and to MC-LF for 30 min at MC concentrations ranging

between 10 nM and 1000 nM. Quantification of MC transport was achieved via immunoblotting and

densitometric greyscale analysis. To obtain Vmax and Km values the Michaelis-Menten equation was

used to fit the data using non-linear regression analysis with GraphPad Prism.

Quantitative real-time PCR

To determine the tissue specific expression of zfOatp1d1 and zfOatp1f1-4, cDNA was synthesized as

described above from adult zebrafish tissues and used as a template for semi-quantitative real time

PCR. A 1:5 dilution of the cDNA was used. 18S-rRNA served as normalization control due to its equal

expression in all tissues (data not shown). Gene specific primers were designed for Oatp1d1 (forward

5´-CACAATCCTCCTGCCAGCAAA-3´, reverse 5´-CCCTATGAAACCACTGACTTGT-3´) and Oatp1f (forward

5´-GGTATAGGAACACTGCTAATGG-3´, reverse 5´-CAGAGGGATAATACTGGCTTC-3´), whereby primer

pairs for zfOatp1f1-4 were 100 % identical for all individual sequences. SensiMix™ SYBR® &

Fluorescein Kit (Bioline, London, UK) was used according to the manufacturer’s instructions. A melting

curve analysis of the products was performed at the end of the PCR reaction to confirm the detection

of a single PCR product. Standard curves for primers used were carried out in n=2-4 independent

replicates, whereby each replicate consisted of technical triplicates. The resulting efficiency of

Page 67: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Material and Methods

- 66 -

amplification was 104±4 % for 18s-rRNA (n=4), 98±1 % for zfOatp1d1 (n=3) and 103±0 % for

zfOatp1f (n=2) (data not shown). Data were analyzed according to the Δct method.

Page 68: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 67 -

RESULTS

Sequence analysis

In order to characterize transport properties of several zfOatps, annotated or predicted zfOatp

sequences found in the refseq database of NCBI (http://www.ncbi.nlm.nih.gov/refseq/) were used as

templates for the amplification of the respective open reading frames, and zfOatp1c1, zfOatp1d1,

zfOatp1f1, zfOatp1f2, zfOatp1f3, zfOatp1f4, zfOatp2b1 and zfOatp4a1 were cloned into the

ptracer™CMV/Bsd plasmid. zfOatp3a1, zfOatp2a1 and zfOatp5a1 could not be cloned for unknown

reasons. The integrity of the cloned zfOatp sequences was validated by sequencing and alignment to

the respective annotated amino acid sequences (TABLE 3, FIGURE 13). The zfOatp1c1 sequence differed

in 5 amino acids in the first 630 amino acids. Furthermore, the cDNA suggested the presence of a splice

variant in that the last exon was spliced in a way that resulted in 5 additional nucleotides thus

encoding a protein of 691 instead of 689 amino acids with a completely different C-terminal end. Thus,

zfOatp1c1 was neglected for further analysis. zfOatp1d1 contained an insertion of 76 additional amino

acids compared to the reference sequence. Sequence variations were observed in all zfOatps with the

exception of zfOatp2b1 (TABLE 3, FIGURE 13), suggesting the presence of single nucleotide

polymorphisms. For zfOatp1f2, zfOatp1f4 and zfOatp4a1 two expression variants were found

(zfOatp1f2-1 and zfOatp1f2-2 respective zfOatp4a1-1 and zfOatp4a1-2), which differed from each

other as well as from the annotated sequence by several amino acids. zfOatp1f2-1 contained a deletion

of 88 amino acids after position 361 compared to the annotated zfOatp1f2, most likely representing a

splice variant with a missing exon.

The membrane topology of the cloned zfOatps was predicted using 3 different algorithms (CHAPTER 4,

MATERIALS AND METHODS). Depending on the algorithm used, the number of predicted TMD varied.

zfOatp1f1, zfOatp1f2-2 and zfOatp1f3, which contained shorter amino acid sequences, were also

predicted to have only 8-10 TMDs rather than the 12 TMDs typically found in mammalian OATPs

(Wang et al., 2008) . Based on multiple alignments, both zfOatp1f1 and zfOatp1f3 were missing the

first TMD while zfOatp1f2-2 lacked at least TMD 8 (SUPPLEMENTAL TABLE 7).

Page 69: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 68 -

Table 3: Variations of cloned amino acid sequences compared to annotated sequences. For detailed amino acid replacement information and accession numbers of the annotated reference sequences see SUPPLEMENTAL FIGURE 5 to SUPPLEMENTAL FIGURE 11.

Protein name

length Variations compared to annotated sequence at aa level

Number of deviating aa aa stretches/gaps other Oatp1c1 691 5 (K7E; G369S; A389T; A426V;

N536S) - Frame difference at aa 632

Oatp1d1 689 2 (R239K; G613D) insertion (76 aa) after aa 377

-

Oatp1f1 561 14 (L40I; I83M; I116T; V183A; S216O; E217Q; I229A; L295V; T353A; I359V; D374E; V386L;

M474L; G499S)

- -

Oatp1f2-1 662 4 (A32C; V87I; F221L; Y325C) - - Oatp1f2-2 574 - Deletion (88 aa) after

aa 362 -

Oatp1f3 588 12 (D59A; F80Y; K85Q; R90S; A295G; S363G; Y447F; S503A; T515N; K516I; N563S; S576C)

- -

Oatp1f4-1 644 9 (I105M; F135I; I320V; W385L; I418V; G427R; T435I;

K512N; L607P)

- -

Oatp1f4-2 644 13 (L82P; I105M; F135I; T136I; T311A; I320V; W385L; L387P;

L416P; I418V;G427R; T435I;K512N)

- -

Oatp2b1 677 - - - Oatp4a1-1 742 4 (M175V; F360S; K365Q;

C531R) - -

Oatp4a1-2 742 3 (F130L; F230S; L512P) - -

Page 70: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 69 -

Fi

gure

13:

Glo

bal

alig

nmen

ts o

f se

vera

l ze

braf

ish

Oat

ps c

lone

d in

thi

s st

udy.

Co

mpa

riso

n to

the

res

pect

ive

anno

tate

d se

quen

ce

reve

aled

seve

ral d

iffer

ence

s in

sing

le a

min

o ac

ids o

r mis

sing

/add

ition

al a

min

o ac

id st

retc

hes.

Amin

o ac

ids w

hich

diff

er fr

om th

e an

nota

ted

one

are

indi

cate

d in

bla

ck in

the

nove

l seq

uenc

e an

d th

e sp

ecifi

c am

ino

acid

exc

hang

e is

dem

onst

rate

d ab

ove.

Red

line

s in

the

sequ

ence

in

dica

te th

at b

oth

of th

e no

vel v

aria

nts c

onta

in th

e id

entic

al a

min

o ac

id w

hich

var

ies f

rom

the

anno

tate

d on

e. T

he O

ATP

supe

rfam

ily si

gnat

ure

(ligh

t gre

en) a

nd K

azal

SCL

21 d

omai

n (d

ark

gree

n) a

re co

nser

ved

dom

ains

. Add

ition

al cr

ucia

l res

idue

s for

tran

spor

ter f

unct

ion

and

loca

lizat

ion

of O

atp1

d1 in

clud

e N

-gly

cosy

latio

n si

tes N

122,

N13

3, N

499

and

N51

2 (p

urpl

e) a

nd H

79 (P

opov

ic e

t al.,

201

3).

Page 71: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 70 -

Substrate specificity

An initial functional screening of the cloned zfOatps was carried out with the known OATP/Oatp

substrates taurocholate (TCA), estrone-3-sulfate (E3S) and methotrexate (MTX) (FIGURE 14).

zfOatp1d1 transported all three substrates. E3S was also transported by zfOatp1f2-2, zfOatp1f3,

zfOatp1f4-1 and zfOatp1f4-2. However, with the exception of zfOatp1d1, none of the other zfOatps

transported TCA or MTX. Furthermore, zfOatp1f1, zfOatp1f2-1, zfOatp2b1 and zfOatp4a1 did not

transport any of the three substrates. Thus, zfOatp1d1 seemed to represent a multi-specific Oatp while

the other transporting zfOatps appeared to be more selective.

Microcystin congener dependent uptake by Oatp

Due to the fact that zfOatp2b1 and zfOatp4a1 did not show transport of OATP/Oatp substrates, these

zfOatps were excluded from further testing with MCs. Although the same holds true for zfOatp1f1 and

zfOatp1f2-1, these zfOatps were included in the initial MC uptake studies as they are part of the

zfOatp1 family for which transport of MCs was expected (Feurstein et al., 2010; Fischer et al., 2010;

Fischer et al., 2005; Komatsu et al., 2007; Lu et al., 2008; Meier-Abt et al., 2007; Monks et al., 2007;

Steiner et al., 2014). All cloned members of the zfOatp1 family were tested for the uptake of different

MC congeners. Therefore, HEK293 cells transiently expressing zfOatp1 subtypes were exposed to

50 nM of MC-LR, MC-LW, MC-RR or MC-LF for 24 h followed by the detection of covalently bound MC-

congeners in western blots with the anti-Adda-antibody (CHAPTER 4, MATERIALS AND METHODS). As

already observed for the OATP/Oatp model substrates, zfOatp1d1 also transported all tested MC-

congeners (FIGURE 15). Although zfOatp1f2-2 and zfOatp1f3 were able to transport E3S they did not

transport any of the four MC-congeners tested. Similar absence of MC transport was observed for

zfOatp1f1. In contrast zfOatp1f2-1 that did not transport TCA, MTX and E3S, indeed did take up MC-LR,

MC-LF and MC-LW. In addition, zfOatp1f4 (variants 1 and 2), which transported E3S, only transported

the MC-congener MC-LF.

Page 72: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 71 -

Figure 14: Transport of Oatp substrates.5 min uptake of 30 nM [3H]TCA, 6.0 nM [3H]MTX and 6.1 nM [3H]E3S in HEK293 cells expressing different zfOatps (black bar, (A)-(G)) was assessed two days after transfection and compared to the uptake into empty vector transfected HEK293 cells (white bar). Significant transport is marked with a star. Columns represent means ± SEM from three independent experiments run in three technical replicates. Statistics: Mann-Whitney test (GraphPad Prism Software Inc.) and p ≤ 0.05 (one-tailed).

Page 73: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 72 -

Figure 15: zfOatp1d1 and zfOatp1f-subtype mediated MC transport. Western blot of MC in zfOatp transfected cells, empty vector (ev) and non-transfected cells (ctrl.), treated for 24 h with 50 nM MC-LR (A), MC-RR (B), MC-LF (C) and MC-LW (D).

Based on the above results, time and MC-congener dependent transport studies were conducted with

zfOatp1d1, zfOatp1f2-1 and zfOatp1f4 (variants 1 and 2). MC-RR exposure of zfOatp1d1 expressing

HEK293 cells resulted in a nearly linear uptake plateauing after 12 h, while a more rapid uptake, linear

up to 6 h, was observed for MC-LR (FIGURE 16). In contrast, transport of MC-LF and MC-LW occurred

within 1 h. A similar MC congener specific uptake was observed in HEK293 cells expressing zfOatp1f2-

1.Thereby MC-LR was taken up much slower than MC-LW and MC-LF. However, transport of MC-LF

and MC-LW by zfOatp1f2-1 was slower than that observed for zfOatp1d1. Comparable transport of

MC-LF to that observed for zfOatp1f2-1 was found for the two zfOatp1f4 variants.

Page 74: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 73 -

Figure 16: Quantification of time dependent MC uptake by zfOatps via densitometry analysis. Data are means + SEM from three independent experiments.

In view of the slower transport velocities for the more hydrophilic MC-LR and MC-RR when compared

to the more lipophilic MC-LF and MC-LW, a more in-depth characterization of the transport kinetics

was considered crucial. For this, MC-congener concentration dependent transport studies were

carried out with zfOatp1d1 and MC congeners MC-LR and MC-LF, whereby the time range for linear

uptake for each of the MC congeners was chosen. MC-LR and MC–LF uptake by zfOatp1d1 expressing

HEK293 cells was saturable with increasing substrate concentrations ranging between 10 and

1000 nM (FIGURE 17). The relative Vmax values calculated based on MC densitometry suggested a 6-fold

higher uptake velocity for MC-LF. The data also suggest a 2-fold higher affinity of zfOatp1d1 for MC-LF

Page 75: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 74 -

compared to MC-LR. It should be noted however, that these kinetic parameters can merely be used for

a comparison within this study and do not lend themselves for a comparison with kinetic values

obtained from experiments using radiolabelled MCs in various transient or stable OATP/Oatp

expression systems.

Figure 17: Concentration dependent uptake of MC-LR and MC-LF (10 nM-1000 nM) mediated by zfOatp1d1 using immunoblotting with anti-ADDA antibody. HEK293-zfOatp1d1 cells were exposed to increasing concentrations of MC-LR for 3 h (A) and MC-LF for 30 min (B), β-actin served as loading control. Quantification of transported MC-LR (C) and MC-LF (D) was carried out using anti-ADDA Westernblot and densitometric analysis. Vmax and Km were obtained by fitting the data from the greyscale analysis to the Michelis-Menten equation. Data are means ± SEM from three independent experiments.

Organ distribution of Oatp1d1 and the Oatp1f subfamily

The organ distribution of zfOatps was determined by semi-quantitative real time PCR and

demonstrated that zfOatp1d1 had a comparably high expression in brain and liver, followed by a

moderate expression in the intestine, a low expression in the kidney and the gills and a marginal

expression in muscle tissue (FIGURE 18). The zfOatp1f subfamily was exclusively expression in the

kidney.

Page 76: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Results

- 75 -

Figure 18: RNA expression of (A) zfOatp1d1 and (B) zfOatp1f using semi-quantitative real time PCR. Expression was normalized to the reference gene 18s-rRNA. Data are mean ± SEM of 3 biological replicates containing the combined organs of 15 fish each.

.

Page 77: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Discussion

- 76 -

DISCUSSION Differences in MC apical toxicity were reported for coregonids and cyprinids, either after

intraperitoneal injection or gavage of specific MC congeners or gavage of bloom material containing

various MC congeners (Ernst et al., 2006; Ernst et al., 2007; Fischer and Dietrich, 2000; Kotak et al.,

1996; Tencalla and Dietrich, 1997). As the toxicodynamic properties of different MC congeners do not

differ dramatically, i.e. a maximum of a factor 2 for PP1 is reported (Fischer et al., 2010; Höger et al.,

2007), it was assumed that the primary factor responsible for the overt differences in apical toxicity

observed between fish species is toxico-kinetics. It was already suggested from earlier studies in

mammalian cells, expression systems and whole animal studies that the overall kinetics depends on

the MC congener to be transported, the capability of a given OATP/Oatp subtype of transporting a

specific MC congener and finally on the OATP subtype expressed in a given organ/tissue. The finding

that the zebrafish zfOatp1d1, zfOatp1f2-1 and both cloned expression variants of Oatp1f4 are capable

of MC-congener transport, whereby the hydrophobic congeners MC-LF and MC-LW were transported

faster in all zfOatp subtypes tested (FIGURE 17), corroborated the MC congener dependent transport

already observed for mammalian OATPs. Moreover, when comparing zfOatp subtypes, differences in

transport velocities for a given MC congener were observed (FIGURE 16), as previously suggested for

the human transporter hOATP1B1 and hOATP1B3 (Fischer et al., 2010). Finally, the cloned zfOatps

demonstrated variant organ expression (FIGURE 18), suggesting that zfOatp1d1 is expressed in

multiple organs, while zfOatp1f1-4, similar to other representatives of the zfOatp1f1 subfamily, is

primarily expressed in the kidney. The expression pattern of zfOatp1d1 observed in our study is

similar to the one found by Popovic et al. (2013). In their study zfOatp1d1 showed highest expression

in liver and brain with about 30 to 50 fold lower expression in skeletal muscle, gills and intestine. Our

results confirm highest expression in liver and brain and 15 to 30 fold lower expression in skeletal

muscle, gills and kidneys. However, intestinal expression was higher than previously reported. Overall,

Oatp1d1 expression in zebrafish differs from the expression pattern of the orthologous Oatp1d1 in

rainbow trout and little skate (Cai et al., 2002; Steiner et al., 2014) where this Oatp appears to be

exclusively expressed in the liver. Obviously the expression of Oatp subtypes, e.g. zfOatp1d1, and their

inherent affinity and capacity of transporting MC congeners can play a major role within the context of

toxicological impact of cyanobacterial blooms. While some organs like the liver and kidney are capable

of regeneration upon MC induced cytotoxicity as long as MC exposure was insufficiently high to

produce mortality, other organs e.g. the brain cannot. Consequently, subchronic and chronic exposure

of zebrafish to cyanobacterial blooms may result in continued accumulation of MC congeners in the

brain, brain pathology and thus possibly behavioral changes in the individual fish, e.g. swimming

activity (Cazenave et al., 2008; Ernst et al., 2007; Tencalla and Dietrich, 1997). In a similar scenario,

rainbow trout or little skate subchronically and chronically exposed to the same cyanobacterial bloom

Page 78: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Discussion

- 77 -

may primarily experience liver and kidney damage followed by regeneration possibly without having

any tangible effect on the behavior or overall survival of the individual or the local population.

Moreover, in a cyanobacterial bloom where primarily MC-LR is present at low concentrations, uptake

of MC-LR via zfOatp1d1 in the gastrointestinal tract would evolve slowly due to the lower affinity and

capacity of zfOatp1d1. Assuming a bloom would primarily contain the more lipophilic MC-LF, rapid

uptake from the gastro-intestinal tract would ensue due to the higher affinity and capacity of

zfOatp1d1 for MC-LF despite the lower presence of zfOatp1d1 in the gastrointestinal tract. Systemic

distribution of MC-LF, if not completely removed via the first-pass effect in the liver, could result in

neurotoxicity and renal toxicity due to the fact that zfOatp1d1, zfOatp1f2-1 and zfOatp1f4 are

expressed in kidney and are well capable of MC-LF transport. Indeed, the scenario described above is

suggested by the findings of Fischer and Dietrich (2000) in carp exposed to oral concentrations of MC-

LR, where time-dependent pathological changes were observed in carp hepato-pancreas and kidney

and detectable MC-LR was observed in brain and muscle tissue at the end of the exposure.

Remarkably, zfOatps cloned in this study demonstrated significant differences in their amino acid

sequence when compared to the annotated isoforms. Whether these differences stem from alternate

splicing or represent single nucleotide polymorphisms (SNPs) is difficult to ascertain with the present

dataset and confirmation on the genomic level would be needed. However, the genetic variations

observed in this study might be due to the fact that sequences of wild-derived zebrafish have been

compared with annotated sequences from inbred strains. Indeed, intrastrain and interstrain variations

are commonly found in zebrafish (Guryev et al., 2006) suggesting that these very likely also occur in

the analyzed Oatp sequences. In human hOATP1B1, hOATP1B3 and hOATP2B1 more than 100 SNPs

were reported (Nies et al., 2013), whereby some of the SNPs may lead to alterations in OATP

expression, localization and/or function (Kalliokoski and Niemi, 2009). Thus despite that SNPs have

not been described in zfOatps so far, the different transport properties of the two zfOatp1f2 expression

variants regarding their transport of MC as well as common OATP substrates might be explained by

their amino acid difference (TABLE 3, FIGURE 13) and thus could support the presence of SNPs.

Furthermore, the other differences observed in the zfOatp sequences e.g. the additional/missing

amino acid stretches in zfOatp1d1/zfOatp1f2-2, most likely result from alternative splicing. A

comparison of zfOatp1d1 with the annotated sequence as well as with the genomic sequence (Acc.#:

NC_007115.6) revealed exon number 9 encoding 65 amino acids to be missing in the annotated

reference sequence at position 375. Furthermore, exon number 10 contains 11 additional amino acids.

This results in a shorter protein, predicted to have only 613 and not 689 amino acids as found for

zfOatp1d1 in this study. Despite the fact that this leads to one TMD less in the annotated reference

sequence, the functionality of this splice variant does not seem to be impaired, as a recent

characterization of this Oatp also showed multi-specific transport (Popovic et al., 2013). Hence, the

Page 79: Molecular cloning, characterization and relevance for

ZEBRAFISH-OATP MEDIATED TRANSPORT OF MICROCYSTIN-CONGENERS

Conclusions

- 78 -

missing TMD in the annotated zfOatp1d1 does not appear to be decisive for the transport function of

the protein. In contrast, a shorter splice variant of rat OATP1B2 (Kakyo et al., 1999) missing one exon

was demonstrated to have a narrower substrate spectrum than the full-length rOATP1B2 (Cattori et

al., 2000; Choudhuri et al., 2000), thereby suggesting that the localization of the missing exon and

resulting missing TMD can be crucial for the transport function. Notably, both Oatp1f-members

containing only 9-10 TMD did not exhibit any transport capabilities for the tested substrates,

suggesting that a minimum of 11 TMD might be required for a functional zfOatp or that the missing

TMD results in aberrant expression of the protein.

CONCLUSIONS In conclusion the data presented confirm that not only zfOatp1d1 but also zfOatp1f2-1 and both

cloned expression variants of zfOatp1f4 are capable of transporting MC-congeners in zebrafish. The

organ-specific expression of zfOatp subtypes and their inherent MC congener transporting capabilities

are key to the apical toxicity observed. Largely depending on the MC congener and concentration

present in a given environmental cyanobacterial bloom scenario, apical toxicity could range from overt

fish mortality, moribund fish, to fish with altered behavioral patterns. The identification of Oatps as

the key parameter for interpreting potential impacts of MCs in fish thus allows for better

understanding of the impact of toxic cyanobacterial blooms on fish populations and the diversity of

aquatic ecosystems.

ACKNOWLEDGEMENTS This work was enabled by financial support from a DFG Center Grant, the International Max Planck

Research School for Organismal Biology (IMPRS), by a Marie Curie International Research Staff

Exchange Scheme Fellowship (PIRSES-GA-2011-295223), and by grants from the National Center for

Research Resources (RR021940) and the National Institute of General Medical Sciences (GM077336

and GM103549) of the National Institutes of Health.

Page 80: Molecular cloning, characterization and relevance for

CHAPTER 5

GENERAL DISCUSSION

Page 81: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 81 -

The Cyanotoxin Microcystin (MC) poses a severe risk to humans and terrestrial animals. But also fish

and other aquatic animals are due to their habitat directly exposed to Cyanotoxins and therefore

especially affected. Consequences of this exposure include reduced fitness, leading to a decline of

whole populations. This might have a huge impact on ecological systems as some species are more

sensitive to MC and thus shifting of the ecological balance can be expected. Moreover, implications for

humans poses a problem as they, as consumers of possibly contaminated fish, might be affected as

well.

A lot is known about the toxicodynamics of MC in fish, however comparably few studies focused on the

kinetics of this toxin, making it difficult to estimate possible adverse effects. Thus, this thesis aimed to

get a better understanding of the kinetics of MC in fish by focusing on one major factor within the

toxicokinetics of MC: the uptake of the toxin into target organs in fish. We therefore wanted to identify

Organic anion transporting polypeptides (Oatps) in different fish species and investigate them

referring their general properties, e.g. expression level in different organs and substrate specificity,

especially of MC-congeners. Our overall hypothesis was that the bioavailability and organ targeting of

MC in fish strongly depends on the expression level and distribution of MC transporting Oatp subtypes.

The major results of this study demonstrate, that at least three different members of fish Oatps exist,

which are able to transport MC: Oatp1d1, Oatp1f2 and Oatp1f4. In the following, their amino acid

identity with Oatps in other species, their organ specific expression in different species, substrate

specificity and uptake properties of different MC-congeners will be discussed and at the same time an

outlook will be given on the research needed to answer unsolved problems. The discussion will focus

on four major topics: in order to draw conclusions about evolutional and functional relationships

between fish transporters and the human and rodent OATPs, a comparison will be conducted referring

amino acid sequence identity and transported substrates in the first and the second part. In the third

and fourth part of MC consequences of intoxication in fish and the role of Oatps within that on the fish

as organism, the ecosystem and humans, will be discussed.

EVOLUTIONARY RELATIONSHIP OF FISH AND MAMMALIAN OATPS Phylogenetic analysis was conducted, in order to classify the novel fish Oatps identified in this study

within so far annotated OATPs/Oatps. Moreover, hypotheses will be formulated regarding genetic

relationship of mammalian and fish OATPs/Oatps based on phylogenetic analysis, functional

properties and organ distribution found in this and other studies.

Phylogenetic analysis of all OATP/Oatp families in mammals and fish revealed that the families OATP1

to OATP5 are present in fish and in mammals whereas OATP6 was not found in fish so far (CHAPER 3).

This suggests that families OATP1 through OATP5 were all present in fish or a common ancestor and

have evolved in mammals though speciation events. The family OATP6 however would have evolved

through gene duplication within the mammalian lineage.

Page 82: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 82 -

Since all so far identified MC transporters belong to the OATP1/Oatp1 family, this family was the focus

of interest. Our results demonstrate, that the novel Oatp1d1 sequences in fish show highest amino acid

identity with fish and mammalian OATP1C1/Oatp1c1 sequences with 47 % to 54 %, whereas the

identity to the mammalian subfamilies OATP1B and OATP1A is slightly less with 42 % to 52 % or 39 to

45 % amino acid identity respectively (SUPPLEMENTAL FIGURE 12). Moreover, phylogenetic analysis in

CHAPTER 2 AND 3 demonstrate that OATP1C1/Oatp1c1 occurs in mammals as well as in fish, while

members of the subfamily OATP1A and OATP1B could only be found in mammals and members of the

Oatp1d as well as Oatp1f subfamily are fish specific. While the novel Oatp1d1 sequences clearly cluster

with other Oatp1d1 from bony fish, the Oatp1d1 sequence from little skate (Leucoraja erinacea) was

also discussed to belong to the Oatp1c subfamily (Popovic et al., 2013). This ambiguous classification

is due to a low amino acid identity with other Oatp1d1 sequences (< 50 %, TABLE 2). Recently it was

proposed however, that the threshold for subfamilies are likely to be less than the original set 60 % ,

when comparing orthologs of species, which are less related than others (Hagenbuch and Stieger,

2013). This might apply to classification of Oatp1d1 in fish as well, since the little skate belongs to the

cartilaginous fish and all other fish presented in the tree to the more recent bony fish. It is tempting to

speculate that Oatp1d1 in the little skate is not only a functional but also a phylogenetic orthologous of

the Oatp1d1 in bony fish, as they are both predominantly expressed in the liver and exhibit

multispecific transport. However, from our phylogenetic analysis including a bootstrap support of

< 50 % a classification as Oatp1d1 cannot clearly be made for the little skate transporter. In fact, this

transporter might belong to a completely new subfamily, which is not annotated yet. For

simplification, this transporter will be referred to as Oatp1d1 in the following discussion. Determining

the phylogenetic relationships between mammalian and fish transporters seems similar ambiguous.

While clustering of the subfamilies within the OATP1/Oatp1 family are supported with a strong

bootstrap value, the phylogenetic position of these clusters is mostly supported with a bootstrap value

below 70 % and therefore the exact position remains elusive (CHAPTER 3). For the human transporter

it was suggested, that OATP1A1, OATP1B1, OATP1B3 and OATP1C1 build a gene cluster as they are all

clustering on chromosome 12. Due to their increasing gene size the authors proposed that OATP1C1,

with the shortest gene represents the most ancestral gene, while OATP1A2, OATP1B1 and OATP1B3

derived through gene duplication (TABLE 4), (Pizzagalli et al., 2002). According to several genome

browsers, the zebrafish (Danio rerio) transporters Oatp1c1 and Oatp1f4 were assigned on

chromosome 4, while Oatp1f1 is located on chromosome 5. The chromosomal location of Oatp1f2,

Oatp1f3 and Oatp1d1 is ambigious. Some genome browsers describe their location as “unplaced

scaffold”, which means a chromosome cannot be assigned, however, all three transporters can be

found on the same scaffold (ZV_NA995), suggesting they are on the same chromosome. Others on the

contrary predict a location on chromosome 4. Assuming that all Oatps from the Oatp1 family are

located on chromosome 4 in fish (except for Oatp1f1), they are likely to build a gene cluster as well,

Page 83: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 83 -

similar to the human transporters. According to the theory that gene length increases through gene

duplication, Oatp1f3 would be encoded by the shortest gene and therefore represents the most

ancestral gene. If that was the case, members of the Oatp1f subfamily should be found in other fish

species as well, which was so far only demonstrated for the fathead minnow (Pimephales promelas,

(Muzzio et al., 2014). Other members of the Oatp1f subfamily as well as Oatp1c1 and Oatp1d1 might

have derived through gene duplication in the fish lineage. If this gene duplication took place before

the fish/mammalian split, it is tempting to speculate that one gene of the mammalian subfamily

OATP1C, OATP1B or respective OATP1A is the ortholog of one gene of the fish subfamily Oatp1c,

Oatp1d or respective Oatp1f. The functional similarity between Oatp1d1 and OATP1B1 respective

OATP1B3 as well as their predominant expression in the liver could be explained by this orthology.

However, functional similarity can not necessarily be related to orthology as already changes in single

amino acids might lead to complete changes of function. Looking at the increasing gene size of fish and

human OATPs/Oatps it is more likely that Oatp1d1 and Oatp1c1 derived from Oatp1f3 through gene

duplication before the fish/mammalian split, as described above, but only OATP1C1 appeared as

ortholog of Oatp1c1 in mammals, while the other Oatp1 members from fish where lost. Other

members of the mammalian OATP1 family would then have derived through gene duplication (as

suggested by (Pizzagalli et al., 2002)) after the fish/mammalian split. In that case, similar function and

location of Oatp1d1 and OATP1B1 respective OATP1B3 would have developed independently in each

lineage.

It becomes obvious, that the phylogenetic relationship between the different OATPs/Oatps in family 1

in these different species cannot be fully resolved in this study. Nevertheless, the theories outlined

above explain how OATPs/Oatps might have evolved.

Page 84: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 84 -

Table 4: Gene lenght and location of human and fish OATPs/Oatps belonging to family 1. In humans all four members are located on chromosome 4. In fish, the chromosomal location is ambiguous and presented according to the respective genome browser. NCBI: National Center for biotechnology Information (ncbi.nlm.nih.gov); ZFIN: ZFIN Gbrowse (zfin.org); Ensembl: (ensemble.org); UCSC: UCSC Genome Browser (genome.ucsc.edu); Vega (vega.sanger.ac.uk). ZV_NA995: scaffold.

Human OATPs

Size [kbp] 1 Chromosomal location1

OATP1A2 66 12

OATP1B1 98 12

OATP1B3 101 12

OATP1C1 58 12

Zebrafish Oatps

Size [kbp] 2 Chromosomal location

NCBI ZFIN Ensembl UCSC Vega

Oatp1c1 52.68 4 4 4 4 4

Oatp1d1 35.87 4 ZV_NA995 ZV_NA995 ZV_NA995 4

Oatp1f1 8.17 5 5 5 5 -

Oatp1f2 11.49 4 ZV_NA995 ZV_NA995 ZV_NA995 -

Oatp1f3 10.22 4 ZV_NA995 ZV_NA995 ZV_NA995 -

Oatp1f4 19.51 4 4 4 4 -

1 (Pizzagalli et al., 2002) 2 according to ZFIN and Ensembl

SUBSTRATE SPECIFICITY AND ORGAN SPECIFIC EXPRESSION OF OATPS In the present study we aimed to identify MC transporting Oatps in fish, which are functional similar to

mammalian OATPs, as we wanted to gain further information on the uptake properties of MC in fish.

Thus, a comparison of the novel fish Oatps and annotated mammalian and fish OATPs/Oatps regarding

their substrate specificity and MC uptake properties is used to draw conclusions about the possible

general function of the transporters identified in the present study.

Uptake of common endogenous and exogenous OATP substrates mediated by the fish transporter

Oatp1d1 and members of the Oatp1f subfamily

Uptake of conjugated and unconjugated bile acids by a sodium-independent system in hepatocytes of

rainbow trout (Oncorhynchus mykiss) and little skate was already suggested more than two decades

ago (Fricker et al., 1994; Fricker et al., 1987; Rabergh et al., 1994). Later, it was demonstrated that

Oatp1d1 is mediating the uptake of the bile acid taurocholate and other substrates (Cai et al., 2002).

Other than that, investigations in fish and especially bony fish remain scarce. Thus, we characterized

fish Oatp transporter belonging to two subfamilies regarding their uptake properties: Oatp1d1 cloned

from rainbow trout and zebrafish as well as the transporters Oatp1f1-4 from zebrafish (CHAPTER 2 AND

Page 85: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 85 -

CHAPTER 4). The transport of three common OATP substrates was tested by overexpressing Oatp1d1

from zebrafish in human embryonic kidney (HEK)293 cells. These substrates, namely estrone-3-

sulfate (E3S), taurocholate (TCA) and methotrexate (MTX), represent steroid hormones, bile acids and

drugs respectively. All of them were transported very efficiently by the zebrafish Oatp1d1 (CHAPTER 2

AND CHAPTER 4). For the rainbow trout Oatp1d1 a comprehensive screening of 14 substrates was

accomplished. Out of those, eight substrates were transported, thus Oatp1d1 exhibits a multispecific

transport. This result is in accordance with results obtained by Popovic et al. (2013), who suggested a

multispecific transport for zebrafish Oatp1d1 using competitive uptake assays with luciferase yellow

in HEK293 cells transfected with the transporter. However it should be noted, that this is not a clear

proof of substrate transport, as the kind of competition was not investigated. Also, Oatp1d1 cloned

from little skate transported a range of similar tritium-labelled substrates (Cai et al., 2002; Meier-Abt

et al., 2007). The rainbow trout Oatp1d1 was further characterized regarding time- and concentration

dependent uptake of E3S, TCA and MTX, which revealed a saturable transport at 13.9 µM, 103 µM and

13.4 µM respectively (CHAPTER 2). In humans, OATP1A2, OATP1B1 and OATP1B3 exhibit a similar

wide range of transported substrates, and the same is true for the rat homologues. Moreover, the

affinities for the three tested substrates are in a similar range (CHAPTER 2, TABLE 1 and references

therein). Therefore, we can conclude that the substrate profile as well as the affinity of the mentioned

three substrates corresponds in great parts with the human OATP1B1 and OATP1B3 but also

OATP1A2 (summarized in (Hagenbuch and Stieger, 2013).

So far, members of the Oatp1f family have not been investigated concerning their substrate specificity.

Here we could demonstrate that at least some members are able to mediate the uptake of E3S,

whereas TCA and MTX were not transported (CHAPTER 4). Due to only little information about the

substrate specificity of members of the Oatp1f subfamily, a comparison with mammalian transporters

based on this knowledge is not possible though. In order to do that, more substrates need to be

screened for their uptake properties and affinities should be compared. In fact, most of the human

OATPs investigated for E3S transport, were able to transport the steroid hormone, although with

different affinities (summarized in (Hagenbuch and Stieger, 2013).

Uptake of MC-congeners mediated by the fish transporter Oatp1d1, Oatp1f2 and Oatp1f4

Our results further demonstrate that the transporters Oatp1d1, Oatp1f2 and Oatp1f4 from zebrafish

and Oatp1d1 from rainbow trout are able to take up various MC-congeners with different velocities.

MC-LF and MC-LW, which are more hydrophobic, were taken up faster compared to the less

hydrophobic MC-LR. MC-RR uptake was comparably small and could only be mediated by Oatp1d1.

Hence we assume that MC-LF and MC-LW are likely to induce higher cytotoxicity compared to MC-LR

while MC-RR mediated cytotoxicity is less compared to MC-LR. This assumption was supported by a

comparison of the kinetics of MC-LR and MC-LW mediated by zebrafish Oatp1d1, which suggested a

Page 86: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 86 -

two times higher affinity for MC-LW (CHAPTER 4). Similar effects were observed by Fischer et al.

(2005) in HEK293 cells transfected with the human transporter OATP1B1 or OATP1B3. Cell viability

assays after exposure with different congeners confirmed a similar toxicity of MC-LW and MC-LF,

whereas toxic effects of MC-LR was 20 (OATP1B1) or 60 (OATP1B3) times less and MC-RR even 500

(OATP1B1) or 300 (OATP1B3) times less compared to these congeners in OATP1B1 or OATP1B3

transfected cells (Fischer et al., 2010). In the same study, inhibition of PP1 and PP2A was similar

between all tested congeners (Fischer et al., 2005). Hence, differences in the cytotoxicity of various

congeners in primary rat hepatocytes and mouse in vivo studies are likely due to variations in their

kinetics rather than their dynamics (Eriksson et al., 1990; Stoner et al., 1989). When comparing the

uptake properties of the rainbow trout Oatp1d1 and human OATP1B3, it seemed like rainbow trout

Oatp1d1 has a lower affinity for MC-LR, compared to the human transporter, because 100 µM TCA

inhibited the uptake of MC-LR in Oatp1d1 transfected cells but not in OATP1B3 transfected cells, even

though both transporters have a similar affinity to TCA (CHAPTER 2, FIGURE 8). This is supported by the

observation, that MC-LR has an EC50 of 257 nM in OATP1B3 transfected HEK293 cells, as

demonstrated with a cytotoxicity assay 48 h after MC-exposure (Fischer et al., 2010). Using the same

parameters, we could not detect a significant reduction in cell viability in Oatp1d1 transfected

HEK293-cells (data not shown). Even though the transfection efficiency might play a role in that as

well, we assume that the Michaelis constant (Km) of MC-LR for Oatp1d1 is in a similar range but

somewhat higher compared to the human OATP1B3. To confirm that, the kinetics of MC-LR has to be

investigated, for that however a quantitative measurement is necessary (e.g. uptake with tritium

labelled MC or HPLC/LCMS).

For the human OATPs the affinity for MC was at the same level or even higher compared to other

substrates like E3S or TCA. The Km value for MC varied between 7 µM (OATP1B1), 9 µM (OATP1B3)

and 20 µM (OATP1A2) (Fischer et al., 2005) while a transport saturation between 10 and 60 µM of E3S

and TCA was reported (summarized in (Hagenbuch and Stieger, 2013). Similar results were obtained

in little skate where Oatp1d1 shows a higher affinity for MC (27 µM, Meier-Abt et al., 2007) than for

E3S or TCA (61/85 µM, Cai et al., 2002). This raises the question whether MC-LR is able to inhibit the

Oatp1d1 mediated uptake of physiological substrates in a competitive manner, which would lead to

several implications for the fish. However, MC-LR did not significantly inhibit the uptake of E3S and

TCA in mice neurons (Feurstein et al., 2010). Also oocytes expressing Oatp1d1 from little skate were

not constrained in their phalloidins transport (Km=2.2 µM) when coincubated with an equal

concentration of MC-LR. A 5-fold higher coincubation of E3S and TCA compared to phalloidin however

resulted in a reduction of phalloidin transport between 50 and 60 % (Meier-Abt et al., 2007). Likewise,

physiological substrates were demonstrated to inhibit the uptake of MC-LR. Oocytes transfected with

OATP1B1 or OATP1B3, exhibited a reduction of almost 100 % in the MC-LR uptake in the presence of

20 fold higher TCA concentrations (Fischer et al., 2005) Similar results were obtained for skate

Page 87: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 87 -

hepatocytes, where 100 μM taurocholate inhibits MC-LR (10 μM) uptake by approximately 50 %

(Runnegar et al., 1999). This leads to the assumption that comparison of the Km values of various

substrates cannot be used to draw conclusions about the adverse effects. A possible reason for that is

the presence of different binding- or translocation sites, which was already shown for the transport of

E3S, mediated by the human transporters OATP1B1 and OATP2B1 and the bovine transporter

OATP1A2. A biphasic kinetic behavior suggests two binding sites, of which one transports the

substrate in a low affinity and the other in a high affinity manner (Liu et al., 2013; Noé et al., 2007;

Shirasaka et al., 2012; Tamai et al., 2001). For fluvastatin however a monophasic kinetic mediated by

the human OATP1B1 was observed, leading to the assumption that this substrate binds to only one of

the putative sites (Noé et al., 2007). Interestingly Noé et al. (2007) demonstrated an inhibition of

fluvastatin and other substrates by the lipid-lowering agent gemfibrozil, while the transport of E3S

was not affected. This kind of substrate specific inhibition was also made for other substances

(Shirasaka et al., 2012; Tirona et al., 2003; Vavricka et al., 2002), which makes a prediction for

interactions very complex. Hence, multiple binding sites might also be present in the fish Oatp1d1,

preventing an inhibition of endogenous substrates. To fully proof that and understand how different

binding sites could influence each other, further analysis regarding kinetics and inhibition studies, but

most importantly structural analysis to find putative binding sites are necessary.

Conclusions referring the possible role of Oatp1d1 and members of the Oatp1f subfamily

Despite the overlapping substrate spectrum of Oatp1d1 with three human transporters (OATP1B1,

OATP1B3 and OATP1A2) and the similarity of MC transport, we suggest that Oatp1d1 in fish has a

functional similar role like the human transporters OATP1B1 and OATP1B3 since they are

predominantly or solely expressed in the liver. Because of this organ specific expression, OATP1B1

and OATP1B3 haven been proposed to play a crucial role in detoxification and hepatic clearance of

albumin-bound amphipathic compounds (Hagenbuch and Meier, 2004). Likewise Oatp1d1 might play

a role in bile salt homeostasis, elimination of steroid hormones and xenobiotica and hormone balance

as previously suggested (Popovic et al., 2013). The uptake of MC might be part of a detoxification

process, which is supported by the biotransformation and elimination via the bile (Kondo et al., 1992;

Kondo et al., 1996; Mohamed and Hussein, 2006; Pflugmacher et al., 1998; Sahin et al., 1996). The

additional expression of Oatp1d1 in the brain moreover suggests a role in transport of hormones

across the blood brain barrier.

Since the transporters Oatp1f1-4 are exclusively expressed in the kidney we propose a major role of

these transporters in kidney mediated toxicity of MC. It is not possible though to clearly identify a

functional similar transporter in mammals. As discussed above, no transporter within the OATP1

family is present in mammals, which is exclusively expressed in the kidney. Although OATP1A2 is

expressed in the kidney, it is additionally expressed in several other tissues, including brain, liver,

Page 88: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 88 -

lung, testes, placenta and prostate (summarized in(Hagenbuch and Stieger, 2013). It also mediates the

transport of several other substrates including TCA, which is not transported by members of the

Oatp1f subfamily. The only other mammalian transporter which was predicted to be kidney-specific is

OATP4C1 found in humans and rats (Mikkaichi et al., 2004). OATP4C1 exhibits narrow substrate

specificity including digoxin, quabain, thyroid hormones, MTX and E3S (Kuo et al., 2012; Mikkaichi et

al., 2004; Yamaguchi et al., 2010). Because it transports thyroid hormones, OATP4C1 was assumed to

be important in the regulation of thyroid hormone concentrations in the proximal tubule cells, as these

hormones play an important role in maintenance of kidney functionality by activation of target gene

transcription (Mikkaichi et al., 2004). In general, the kidney is responsible for homeostasis of

endogenous and exogenous substances via tubular secretion and reabsorption (Davies et al., 2001).

This is partly mediated by transporter of the ATP-binding cassette (ABC) superfamily and the solute

carrier (SLC) family, including OATPs. The direction of the transport (secretion or reabsorption)

mainly depends on the specific location of the transporter in the polarized cell, i.e. on which

membrane it is expressed. In rats, OATP4C1 is expressed at the basolateral membrane, while rat

OATP1A1 and OATP1A3, as well as human OATP1A2 are expressed at the apical membrane (Bergwerk

et al., 1996; Lee et al., 2005; Masuda et al., 1997). Hence, a similar function in reabsorption or

secretion might be subscribed to Oatp1f1-4, to better understand the exact role however, a further

screening of other substrates and the identification of the exact localization might be advantageous. As

Oatp1f1-4 are fish specific, they might fulfil functions in fish which are covered by several transporters

in mammals.

Possible implications of the structure (SNPs, splice variants) on the function

Notably, different members of the Oatp1f subfamily exhibit differences in substrate uptake despite

their amino acid identity between 84-90 % (SUPPLEMENTAL FIGURE 12). While Oatp1f1 and Oatp1f3

neither transported E3S nor MC, Oat1f2 and Oatp1f4 demonstrated an uptake of both. Therefore the

difference between their sequences seems to be crucial for functionality of the transporter. Even

deviations within variants of one member led to differences in their uptake properties. The most

prominent example for these differences are the two variants of Oatp1f2. While Oatp1f2-1 transported

E3S, it did not transport MC (CHAPTER 4). The opposite effect was observed for Oatp1f2-2. Oatp1f2-1

varied in four amino acids compared to the annotated Oatp1f2 whereas Oatp1f2-2 revealed a deletion

of 88 amino acids resulting in one transmembrane domain less and a deprivation of 3 cysteine

residues in the Kazal domain. Hence, we assume that the exchange of one or more amino acid is crucial

for the MC transport while the deleted sequence contain functional sites for E3S transport. We assume

that alterations of single amino acids are likely caused by single nucleotide polymorphisms (SNP) and

deletions and inserts are the result of alternative splicing. SNPs are mainly contributing to variations

in the genome of an organism (Sachidanandam et al., 2001) and have indeed been described in several

Page 89: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 89 -

OATP transporter as well (Nakanishi and Tamai, 2012; Nies et al., 2013). The impact of SNPs can be

tremendous as it can lead to altered gene expression, protein localization and function as summarized

in (Nakanishi and Tamai, 2012). This was recently also shown by Huang et al. (2013): the exchange of

one tryptophan residue in OATP1B3 resulted in a structural alteration accompanied by a conversion of

two binding sites into one.

Likewise, alternative splicing has also been observed in OATPs. Alternative splicing is a regulating

process which occurs after DNA has been described into messenger RNA (mRNA). Specific exons are

included or excluded from the final mRNA, thereby resulting in variations of the final protein. This

might lead to differences in localization and functionality as shown for the human OATP3A1, where

two splice variants were isolated whereby one was lacking 18 amino acids on the c-terminal end.

While both exhibited similar broad substrate specificity, one splice variant was predominantly

expressed in testis and brain whereas the other exhibited a wide tissue distribution (Huber et al.,

2007). Also in the mouse OATP1A4, where three isoforms have been described, a differential tissue

distribution and similar but not identical transport properties were observed (Ogura et al., 2001).

Therefore, it is conceivable that the missing amino acids in Oatp1f2-2 led to alterations in the E3S

uptake. Additionally, the missing stretch contained 3 Cysteine residues of the Kazal domain. Eight out

of ten conserved cysteine residues are part of this domain. Due to their conservation between

OATPs/Oatps they have been suggested to be important for transport function (Hänggi et al., 2006).

IMPACT OF MC ACCUMULATION AND OATP EXPRESSION IN FISH: RELEVANCE FOR FISH

Relevance for the individual

The consequences of MC accumulation that might be expected for the individual fish are toxicosis of

target organs and subsequent pathologies, as well as additional implications due to reduced uptake of

endogenous substrates of Oatps. We demonstrated that Oatp1d1, which has been found in all fish

species investigated so far, is mainly expressed in the liver and to a minor extent in the brain. This was

the case for rainbow trout, whitefish (Coregonus wartmanni), common carp (Cyprinus carpio, CHAPTER

2 AND 3). The same organ specific expression was found in the little skate (Cai et al., 2002). In common

carp, we additionally found an expression of Oatp1d1 in the kidney. Moreover, zebrafish Oatp1d1

exhibited a comparably ubiquitous expression as RNA levels in liver and brain were not statistically

different from each other. An additional expression of Oatp1d1 mRNA was discovered in the intestine

and gills, even though to a minor extend (CHAPTER 4). Oatp1f2 and Oatp1f4 from zebrafish, which also

transported MC, were specifically expressed in the kidney (CHAPTER 4). These results are

corresponding to the organ specific MC accumulation and pathologies found in fish including common

carp, rainbow trout and whitefish (Ernst et al., 2007; Fischer and Dietrich, 2000; Moutou et al., 2012;

Papadimitriou et al., 2013; Råbergh et al., 1991; Singh and Asthana, 2014; Tencalla and Dietrich,

Page 90: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 90 -

1997; Tencalla et al., 1994). As we could demonstrate that Oatp1d1 is transporting MC in rainbow

trout and zebrafish and Oatp1f2 and Oatp1f4 are additional MC-transporters in zebrafish, we conclude

that these transporters play a major role in the toxicokinetics and are responsible for the

organotropism of MC meditated toxicity.

The liver as target for MC-uptake

The liver was suggested to be the main target of MC as the highest amounts of MC have been found

here shortly after the intoxication with MC (Ernst et al., 2006; Fischer and Dietrich, 2000; Tencalla

and Dietrich, 1997; Williams et al., 1997; Williams et al., 1995). We could confirm in the present study

that this can be explained by the dominant expression of the MC transporter Oatp1d1. Nevertheless,

other hepatic zebrafish transporters, like Oatp2a1 and Oatp2b1, whose MC uptake properties have not

been investigated yet, might contribute to that MC uptake. As the liver is the detoxifying organ of the

body it seems likely that Oatp1d1 plays a key role in the hepatic clearance of the toxin. This is

supported by the fact that biotransformation into MC/GSH and MC/CYS conjugates and a biliary

excretion and therefore elimination of the toxin takes place in the liver (Pflugmacher et al., 1998;

Sahin et al., 1996). The expression of this transporter in the liver would therefore be beneficial for the

species since it can reduce the bioavailability of the toxin in the body. Hence a higher expression and

affinity for MC would have been a selective advantage in evolution. Nevertheless, pathological

alterations upon MC exposure were found after giving relevant doses of MC to fish: Maximum amounts

of MC can be up to 25,000 µg/L of the dissolved toxin the water and up to 7300µg/g dw in the

cyanobacterial cells (Sivonen and Jones, 1999). In force feeding experiments, concentrations as low as

1700 µg/kg body weight (bw) were given to carp, which resulted in pathological alterations of several

organs (Tencalla, 1995). Ernst et al. (2007) exposed whitefish (Coregonus lavareutus) to relevant low

(0.3 µg/L), medium (1.8 µg/L) and high doses (11 µg/L) of Planktothrix cells, containing MC-

equivalent (equ.) in the water and observed histopathological changes in the liver in a time- and dose-

dependent manner. Also in silver carp (Hypophtalmichthys molitrix), smelt (Osmerus eperlanus), perch

(Perca fluviatilis) and ruffe (Gymnocephalus cernuus), taken during blooms from Lake Taihu in China or

Lake IJessel in the Netherlands, abnormalities in liver and kidney were found including inflammation,

degeneration and necrosis (Ibelings and Havens, 2008; Li et al., 2007). These results lead to the

question whether the liver can just not keep up with biotransformation and elimination at high

concentrations or if the uptake of MC is rather a side effect of Oatp transporting other substrates. It is

likely that MC is involuntarily transported by Oatps, as it exhibits structural similarity to physiological

substrates, mainly other peptides (e.g. enkephalins, phalloidin, vasopressin, deltorphin). This might

explain why MC can be transported in tissues, where its accumulation can only be a disadvantage like

the muscle or the brain. Also, in case of a carrier mediated uptake via Oatps in the intestine, as

Page 91: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 91 -

suggested for zebrafish (CHAPTER 4), the benefit remains unclear as this is the entrance into the portal

vein in the first place.

The kidney as target for MC-uptake

The active transport of MC into the kidney is possibly primarily mediated by Oatp1f2 and Oatp1f4 in

zebrafish. The expression of this transporter in other fish has only been shown in the fathead minnow,

where the expression was also kidney specific (Muzzio et al., 2014). In all other fish species these Oatp

subtypes need to be identified and characterized to display their importance in MC uptake in the

kidney. At least in common carp, Oatp1d1 could be partly responsible for this uptake as well, since this

transporter could also be detected in this fish species on the RNA level (CHAPTER 3). Moreover, just as

in the liver, the ubiquitous subtypes Oatp2b1 and Oatp2a1 are expressed in even higher amounts in

the kidney compared to members of the Oatp1f subfamily in zebrafish (Popovic et al., 2010). Their

ability to transport MC however has not been investigated yet. The accumulation of MC in the kidney

might be due to a second way of excretion of the toxin. This has indeed been shown in mice, where 9 %

of MC was found in the urine (Robinson et al., 1991). Although, this kind of evidence is missing in fish,

the fact that conjugated forms of MC have mainly been found in the kidney supports the idea of

elimination through the kidney (He et al., 2012; Ito et al., 2002; Li et al., 2014a). However, to

understand the exact toxicokinetics of MC in the kidney, several questions need to be answered: it has

not been investigated yet whether MC if filtrated through the glomerulus or if it is taken up actively in

the tubular lumen and is therefore reabsorbed. The fact that MC, as well as its conjugated forms, are

rather small (< 2 kDa) argues for a filtration, as substances up to 10 kDa can pass the filtering

membrane in the glomerulus (Davies et al., 2001). When MC is bound to proteins though, it will not be

filtrated by the glomerulus and remains in the blood for active uptake. In fish it was shown, that about

4 % of the MC is bound to proteins in the plasma (Zhang et al., 2013). Additionally, it needs to be

investigated whether conjugated forms of MC can be transported by Oatps and in which direction this

transport is taking place. In humans, OATP1A2 is expressed at the brush boarder of the tubular cells

and is most likely bidirectional (Hagenbuch and Meier, 2004), while OATP4C1 is expressed at the

basolateral membrane, probably causing an influx transport. An influx transport at the basolateral

membrane as well as an efflux transport from the tubular cells into the tubular lumen would suggest a

secretion of substrates from the capillaries into the tubular fluid, which eventually will be excreted as

urine. A transport directed in the other direction however could support reabsorption of MC and its

conjugated forms from the tubular lumen back into the tubular cells or even the blood system,

whereby an accumulation in tubular cells would occur or the bioavailability of MC in the blood would

increase. The expression of Oatp1f2 and Oatp1f4, the so far only known MC transporters in the kidney

of fish, needs to be investigated on a protein level to better understand these mechanisms. If

Page 92: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 92 -

antibodies were available for these transporters it would be interesting to stain kidney slices of

zebrafish to ascertain their exact localization of the membrane.

The brain as target for MC-uptake

Subsequent pathologies to an MC accumulation in the brain remain elusive in fish. Although the

expression of Oatp1d1 needs to be confirmed on a protein level as well, this transporter might be

responsible for the accumulation found in fish brain (Fischer and Dietrich, 2000; Moutou et al., 2012;

Papadimitriou et al., 2012). Whether this expression however leads to subsequent pathologies needs

to be investigated. Alterations in swimming activity might be a hint for neurotoxicology, as suggested

by (Cazenave et al., 2008). It remains unclear though, whether this is actually caused by a neuronal

impairment or if this effect is rather due to general reduction of fitness and organ pathologies. In

humans though, neurotoxicology was observed when 131 hemodialysis patients were accidental

intravenously exposed to MC via dialysis water in Caruaru (Brazil). 89 % exhibited acute symptoms of

neurotoxicity including deafness, tinnitus and visual disturbance (Carmichael, 2001; Soares et al.,

2006). Additionally, recently impairment of spatial learning and memory accompanied with enhanced

activation of astrocytes could be observed in rats (Li et al., 2014b). In neuronal and whole brain cells

of mice, cytotoxicity upon MC exposure of various congeners could be observed (Feurstein et al., 2009;

Feurstein et al., 2010). The expression of the MC transporter OATP1A2 in those cells might be

responsible for the uptake across the human blood brain barrier. Due to these observations in

mammals it is likely that pathologies in fish brain can be expected as well, however direct proof of that

still has to be found.

Other organs as targets for MC-uptake - presence of additional transporters?

Accumulations of MC have also been found in gut, gills and muscle, however subsequent pathologies

were only found in the gut (Carbis et al., 1996; Fischer and Dietrich, 2000; Freitas de Magalhães et al.,

2001; Poste et al., 2011; Sahin et al., 1996; Song et al., 2007). Of the investigated species in this study,

only in zebrafish an Oatp1d1 expression was found in the gastrointestinal tract and could explain

pathology in this organ (CHAPTER 4). However, the studies mentioned above, were conducted with

rainbow trout or carp. Since no or vanishingly low expression of Oatp1d1 could be found in gut, gills

and muscle in those species, either paracellular transport of MC takes place or other Oatp transporters

are expressed in these organs, which also are able to transport MC. In humans, besides the liver

specific transporter OATP1B1 and OATP1B3, only OATP1A2 was demonstrated to transport MC

(Fischer et al., 2005). OATP1A2 is a ubiquitous expressed transporter, including brain, kidney, liver,

lung, testes, placenta and prostate (Roth et al., 2012). The only human transporter, which so far has

been found to be expressed in the intestine and in the skeletal muscle is OATP2B1 (summarized

in(Hagenbuch and Stieger, 2013). This transporter however was not found to transport MC (Fischer et

al., 2005). In zebrafish, we additionally identified Oatp1f2 and Oatp1f4 as MC transporters. Since

Page 93: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 93 -

Oatp1f2 and Oatp1f4 are solely expressed in the kidney, as shown in our and other studies (Muzzio et

al., 2014; Popovic et al., 2010), this also does not explain the accumulation of MC in gut, muscle and

gills. An identification of these transporters in rainbow trout and carp with subsequent

characterization regarding MC uptake would help to resolve the question whether a similar function

and organ expression is exhibited. A screening of other fish Oatp transporters for MC transport would

help resolving this question. According to Popovic et al. (2010), Oatp3a1 and Oatp2a1 are ubiquitous

expressed transporters in the zebrafish, including gills, intestine, muscle and kidney and they are also

expressed in similar amounts like Oatp1d1. Hence they might be interesting candidates to test for MC

transport. It should also be noted, that transporters belonging to the ABC superfamily are likely to play

a role in MC transport as well. As second family next to the solute carriers, they play an important role

in drug disposition but also transport of endogenous substrates. Thereby the specific expression of

ABC- and SLC transporters on polarized cells often leads to a directed transport of substrates, which

makes both transporters important for the uptake and elimination of substances like MC (You and

Morris, 2007) Hence, it would be very interesting to screen ABC-transporters (e.g. multidrug resistant

proteins (MDR) or multidrug resistance associated proteins (MRPs) regarding their ability to

transport MC in fish.

Adverse effects between MC and other Oatp substrates

Due to uptake of MC, the transport of endogenous Oatp substrates might be reduced. However,

according to the present state of knowledge, we do not expect alterations due to reduced uptake of

endogenous substrates of MC. As discussed above, MC usually does not seem to inhibit the uptake of

other substrates, including TCA, E3S and Phalloidins (Feurstein et al., 2010; Meier-Abt et al., 2007). An

alteration in thyroid hormone levels upon MC exposure has been suggested recently, however rather a

decrease than an increase in the blood was observed (Yan et al., 2012). If MC was competing for

uptake via Oatp with thyroid hormones, one would rather expect a higher thyroid concentration in the

blood than a reduced one. The other way around though, the fact that MC transport was inhibited by

typical Oatp substrates, might offer possibilities to prevent MC induced pathologies by administration

of other Oatp substrates. This could already be shown for mice, which were protected of lethal

intoxications with MC when the antibiotic rifampicin, a known Oatp substrate, was administered

simultaneously (Hermansky et al., 1991).

Relevance for the ecological balance

The ecological balance might be impaired due to MC-accumulation, as various species are more

affected, leading to a reduced fitness or even death of animals. We demonstrated the existence of

Oatp1d1 in several fish species, including rainbow trout, whitefish, common carp and zebrafish. We

observed species specific differences in the organotropism of the Oatp RNA expression level as well as

Page 94: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 94 -

variations in the relation of expression level in different organs (CHAPTER 3). We therefore propose

that the observed species specific differences in MC accumulation, MC-targeting and subsequent

pathologies can partly be explained by the expression level and distribution of Oatp1d1.

The main uptake route of MC was suggested to be feeding with subsequent uptake via the digestive

system, rather than uptake of dissolved MC through the gills (Ibelings and Havens, 2008). Hence, it

seems conspicuous that herbivorous and planktivorous fish, which filter food are more likely to take

up cyanotoxins than omnivourous and carnivorous fish. However, accumulations of MC have been

found in species of different trophic levels (Chen et al., 2009; Gkelis et al., 2006; Schmidt et al., 2013;

Singh and Asthana, 2014). Even though MC amounts varied in fish, those studies do not provide a

perceptible tendency that fish of a specific trophic level are more susceptible to MC incorporation.

Indeed, even fish of higher trophic levels are threatened, as shown for the pumpkinseed sunfish

(Lepomis gibbosus), which was fed with MC containing zooplankton. 80 % of the non-covalently bound

MC in the zooplankton was transferred to the sunfish (Smith and Haney, 2006). Even though it is still

debated whether biodilution or bioaccumulation of MC takes place, a transfer of MC through the food

chain seems unquestionable (Kozlowsky-Suzuki et al., 2012).

This raises the question whether species susceptibility towards MC really exists or if species just have

to deal with different amounts of toxins due to their diet and habitat.

When rainbow trout and common carp have been directly gavaged with the same amount of MC

(1700 µg/kg bw) 100 % mortality was observed in carp, while no rainbow trout died. Rainbow trout

mortality occurred at only 6600 µg/kg bw (Tencalla, 1995; Tencalla et al., 1994). As common carp is

omnivorous and obligatory herbivorous while rainbow trout is carnivorous, they exhibit differences in

their gastrointestinal tract (GIT). Rainbow trout have a short, thick walled gut while cyprinids have a

comparably long intestine and therefor higher resorptive capacities (Kapoor et al., 1976). This larger

surface area and higher absorptive capacities might lead to more uptake of MC per time unit into the

blood, as discussed before (Fischer and Dietrich, 2000).

On the contrary, it was found that the amount of MC in the GIT not necessarily correlates with the

amount in the liver. This was observed by Chen et al. (2009), who quantified the amount of MC in fish

species with different diets. While the phytoplanktivorous silver carp (Hypophtalmichthys molitrix)

contained 235 µg/g dw MC-LRequ. in its intestine, which was almost as much as found in the scum, the

other fish species, including the omnivour goldfish (Carrasius auratus), the omnivorous common carp

(Cyprinus carpio) and the carnivorous Culter ilishaeformis, contained no more than 100 ng/g dw MC-

LRequ. in their intestine. Nevertheless, the MC content in the liver was similar, and even highest in

omnivorous goldfish with 150 ng/g dw compared to 60 ng/g dw in the silver carp.

Page 95: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 95 -

These results suggest that species feeding on plankton might be less sensitive towards MC, despite the

fact that they are exposed to higher concentrations and even have a higher bioavailability in the blood.

A possible explanation is the amount of MC transported into the target organs, and therefore Oatp

expression level and -profile as well as affinity for MC. Moreover, duration and amount of

biotransformation and elimination are likely to play a role. These factors might have been adapted

over the time, as they would be beneficial for species being exposed to high MC concentrations.

Interestingly, despite their limited possibility to avoid the bloom, it seems like fish in general have a

higher tolerance towards MC compared to mammals. This was mostly assigned to different profiles in

toxicokinetics. Rainbow trout and carp withstood up to 10 times higher concentrations which were

already lethal in mice (Botes et al., 1984; Carmichael et al., 1988; Kotak et al., 1996; Råbergh et al.,

1991; Tencalla et al., 1994). Several reasons for that were discussed, including a faster MC

accumulation in the liver of mice than that of fish (as confirmed with radioactive labeled MC (Brooks

and Codd, 1987; Williams et al., 1995), a more rapid clearance of MC in salmonids compared to mice

(since after 8 days the liver and carcass still retained 50 % in mice and only 8 % in salmon (Robinson

et al., 1991; Williams et al., 1995) and a reduced binding to plasma proteins in fish compared to

mammals and therefore a shorter time of bioavailability (Zhang et al., 2013). A resistance of organisms

being exposed to cyanotoxins has also been observed in the planktonic crustacean Daphnia galeata.

When resting Daphnia-eggs have been hatched, it was observed that those formed during high

densities of cyanobacteria in Lake Constance are more resistant to MC exposure, while those eggs

formed in times with low amounts of cyanobacteria exhibited greater inhibition of growth (Hairston et

al., 1999).

To test whether a species specific adaption within different fish species is possible, it would be

interesting to know if LD50 values exhibit differences when the same amount of MC is bioavailable.

This could be investigated using ip injection as exposure route. So far, only few MC concentrations

which lead to mortalities have been tested and can be used as approximate value for assessing the

LD50: 100 % mortality was found 26 h after applying 1000 µg/kg MC-LR to rainbow trout, while 400 µg

MC-LR/kg bw did not result in any dead animals (Kotak et al., 1996). In another study, 100 %

mortality was observed in rainbow trouts after 24 h when applying 550 µg MC-LR/kg bw (Tencalla et

al., 1994). In carp a concentration between 130 and 300 µg MC-LR/kg bw did not lead to mortalities

while at 550 µg MC-LR/kg bw carp exhibited a severe and morbid state within 24 h (Råbergh et al.,

1991). This suggests rather similar sensitivities of MC when similar amounts are in the blood. Carbis et

al. (1996) however found all investigated carp to be dead after 5 h when exposing them ip to 50 µg/kg

bw. It should be noted however that this concentration represents MC equivalents and is not congener

specific. Therefore other, more toxic congeners might have been involved. However a detailed

examination of LD50 values is missing. To test whether different fish species exhibit variations in their

Page 96: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 96 -

MC sensitivity, it would be necessary to parallel expose different fish species to several concentrations

of a specific MC-congener. Through investigation of the LD50 value and observation of the pathologies

in the relevant organs (liver and kidney) a more profound statement could be made referring species

specific sensitivity. To understand the reasons for that, Oatp expression levels, amount of the free and

protein bound toxin and its conjugates should be measured in the target organs and in the blood, urine

and bile to include the impact of uptake, biotransformation and elimination. Additionally it would be

interesting to compare MC-affinities of Oatps from different species in order to investigate whether the

uptake capacity differs as well.

A direct comparison of the Oatp1d1 RNA expression level between different fish species using an

equivalent expressed reference gene may provide insights as to whether a higher Oatp expression in

one species can be expected. As we were interested in the organ distribution of Oatp1d1 RNA within

one species, we focused on choosing reference genes being equally expressed within those organs

rather than across different species. Therefore our data does not allow comparison of the expression

level of Oatp1d1 RNA across different species. Nevertheless, we can see differences of the

organotropism when comparing Oatp1d1 RNA levels in different species. The relation between brain

and liver showed variations in the species investigated and might give a hint that also expression

levels alter. The most prominent difference we observed was the additional expression of Oatp1d1 in

the kidney of carp, which was not present in the other investigated species. Whether this poses a

higher risk to carp due to a higher intoxication of the kidney or if this is even beneficial because of a

higher clearance rate remains unclear, as the exact role of the kidney in MC elimination is not

understood.

IMPACT OF MC ACCUMULATION AND OATPS IN FISH: RELEVANCE FOR HUMANS Accumulation of MC in fish tissue might have consequences for humans as well, as not only economic

losses in aquaculture but also implications for human health can be expected.

Economic losses

Especially in aquaculture, cyanobacterial occurrence poses a problem, as intoxication of aquaculture

species or prey with MC leads to tumors or diminished health like reduced nutrition and growth rate

or incudes stress response. Hence the product cannot be sold anymore or additional costs arise due to

increased maintenance like chemical and physical treatment or additional food and time needed

(Smith et al., 2008). That aquaculture systems are affected has been shown in various studies, which

observed toxicosis and mortality of pond farming catfish (Pelteobagrus fulvidraco)and white shrimp as

well as net-pen farming Atlantic salmon (Salmo salar, (Andersen et al., 1993; Kent et al., 1996; Zimba

et al., 2001; Zimba et al., 2006).

Page 97: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 97 -

Not only the direct exposure of fish to MC or MC containing cells appears to be a problem in

aquaculture but also trophic transfer of MC was demonstrated within various species. In Atlantic

salmon, hepatotoxins were transferred through toxin containing, natural biota and resulted in net-pen

liver disease (Andersen et al., 1993; Kent et al., 1996). Trophic transfer was shown from zooplankton

to fish but also aquatic plants and macroalgae as well as other animals like snails, crayfish, fish, and

shrimp represent vectors (Engström-Öst et al., 2002; Karjalainen et al., 2005; Pflugmacher et al.,

1999; Smith and Haney, 2006). Therefore the economic productivity of the aquaculture is also

reduced indirectly as lower trophic levels are affected as well. However, as discussed above, fish are

probably more tolerant to MC due to their coevolutionary adaption. On the one hand, this is

advantageous for aquaculture as economic losses are reduced, but on the other hand it might increase

the risk for human exposure through fish consumption, since trophic transfer is likely to affect humans

as final consumer of fish from aquaculture as well as lakes. (Smith et al., 2008).

Consequences for human health

Since the transfer of MC through the food chain could be shown, MC-contaminated fish might also pose

a threat to human consumers. Mainly, accumulation of MC in muscle tissue poses a threat to humans as

this is usually the edible part of the fish. Accumulations have often been shown to be above the

tolerable daily intake (TDI) value recommended by the World Health Organization (WHO), hence the

consumption of fish was proposed to be not safe for humans (Freitas de Magalhães et al., 2001; Poste

et al., 2011; Song et al., 2007). According to the WHO a TDI of 1 µg/l in drinking water was suggested

for MC-LR, furthermore a TDI of 0.04 µg MC-LR/kg bw was proposed (Kuiper-Goodman et al., 1999;

WHO, 1998). The interim maximum acceptable concentration (IMAC) in fish used for human

consumption was determined as (Falconer, 2001):

𝐼𝐼𝐼𝐼 =𝑇𝑇𝐼 × 𝐵𝐵 × 𝑃𝑃𝑇

𝐼𝐴𝐼

BW represents the average human body weight, POT the proportion of toxin consumed in form of

contaminated fish and AFC the average fish consumption. In those studies, calculations where based

on a person with an average weight of 60 kg, consuming 100 g fish/day. The proportion of toxin was

assumed to be 1, as the risk assessment was solely based on fish as vector for the toxin. Thereby an

IMAC of 24 µg/kg wet weight (ww) was calculated. The MC concentrations found in fish muscles

exceeded this value by about ~80 fold, as they were up to ~1900 µg/kg ww in muscle tissue (Poste et

al., 2011).

Nevertheless, we suggest that calculation of the TDI value and consequently recommendations not to

eat fish above specific MC content in the muscle should be more critical. First of all, results are

ambiguous, as other studies could not find MC in muscle tissue (Deblois et al., 2008; Deblois et al.,

Page 98: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 98 -

2011; Kohoutek et al., 2010; Lei et al., 2008; Palikova et al., 2011). This is likely because of

differences in experimental performing, but could also be due to differences in the species

investigated. As MC distribution is species specific, probably due to the Oatp expression profile as

discussed above, it would be advantageous to include the species specific effect. Secondly, the

recommended TDI value is based on toxicity of MC-LR (Kuiper-Goodman et al., 1999). However,

analyses were conducted using ELISA (Peng et al., 2010; Poste et al., 2011), which cannot discriminate

between congeners, therefore it is likely that the observed accumulation does not represent the MC-LR

congener, but rather MC-RR. This is supported by several studies which discriminated between the

congeners using other methods like HPLC or LCMS. Zhang et al. (2009b) found all investigated

congeners in liver tissue of silver carp, whereas only MC-RR was found in muscle tissue. Similarly

under laboratory conditions only MC-RR but not MC-LR could be found in silver carp muscle (Xie et al.,

2005). The toxicity of MC-RR however has suggested being much lower than MC-LR, as shown with the

respective LD50 values in mice (Sivonen and Jones, 1999). In humans it could be shown that MC-RR is

taken up by OATP1B3 with a lower uptake rate compared to MC-LR (Fischer et al., 2010). The same

was demonstrated for zebrafish Oatp1d1 in this study. Therefore accumulation of this congener is also

likely to occur in a less sever way than MC-LR.

The predominant accumulation of MC-RR can be explained in two ways: (i) Either MC-LR cannot enter

the muscle cells. This is in accordance with our findings, which showed no Oatp1d1 expression in the

investigated species. At the same time, the less hydrophobic MC-RR might be able to at least partly

cross membranes passively or is alternatively taken up by Oatp transporters, for which MC transport

has not been investigated so far. (ii) Alternatively MC-LR enters the muscle cells with an Oatp

mediated transport, however those Oatps are not characterized for their MC transport yet. The reason,

why this congener is rarely found in muscle tissue can be explained by methodical restrictions, e.g.

purification of MC derived from the tissue for measuring the amount. Indeed, only recently it was

demonstrated, that the recovery rate of MC from human blood depends on the surface adsorption of

aromatic moieties to lab ware and hydrophobicity of the congener (Heussner et al., 2014). As MC-RR is

less hydrophobic than MC-LR and exhibits a lower adhesion capability, it is likely that also in the

purification of MC from fish tissue a better recovery from MC-RR can be expected.

To approach a profound risk assessment, we propose that several other factors should be taken into

consideration. The TDI should be based on the amount of congeners actually present in the fish tissue,

whereby their specific toxicity for mammals should be considered. Since quantification of MC-amounts

and discriminating the congeners still need improvement, we propose that knowing the expression

profile of different Oatp suntypes and their ability to transport different congeners might help to

estimate the likelihood of various congeners taken up into edible tissue. Thereby it would be

advantageous to specify this approach for the fish species and its specific organ distribution.

Page 99: Molecular cloning, characterization and relevance for

GENERAL DISCUSSION

- 99 -

FINAL CONCLUSIONS Taken together, we could isolate and characterize Oatps from fish, which are also able to transport MC-

congeners. With that, we demonstrated the existence of Oatps in bony fish, which are in parts

functional similar to human transporters. These transporters give further insight into the

toxicokinetics of MC in fish and help to understand how and to what extend MCs target specific organs.

However, a further screening of other transporters is necessary, as it remains unsolved how MC can

reach organs which do not express the novel transporter, identified in this study. Furthermore, the

expression level in different organs depends on the species investigated, which could be reason for

species specific sensitivity towards MC. This can have direct consequences on the balance of an

ecosystem. With the identification of Oatp1d1, Oatp1f2 and Oatp1f4 we finally showed that several

transporters are involved in the uptake of MC, whereby the localization of these transporters can give

information about the targets of MC. Different MC congeners are transported with varying capacity,

suggesting that congeners with a lower uptake rate are also less toxic for the fish. Even though more

investigations are inalienable to fully understand the toxicokinetics of MC in fish and possible

implications of MC accumulation on individual fish, the ecosystem and humans, the present study

provides further understanding of the toxicokinetics of MC in bony fish and showed that the

bioavailability and organ targeting of MC strongly depends on the expression level and distribution of

several Oatps in fish. This information can help to estimate the risk of MC accumulations for fish, the

ecological system and also humans.

Page 100: Molecular cloning, characterization and relevance for

CHAPTER 6

SUPPLEMENTAL MATERIAL

Page 101: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 103 -

SUPPLEMENTAL FIGURES

Supplemental Figure 1: For use in qRTPCR primer quality for 18srRNA (A+E), Ef1α (B+F) and rtOatp1d1 (C+G) was tested by standard curves (A-C) and melting curves (E-G). (A-C) A 1:5 dilution series of cDNA was used as template for the qRTPCR and respective threshold values are shown as a function of the logarithmized dilution values. From the consequent linear equation the amplification efficiency (E) can be calculated. Shown is the mean ± SD of three determinations of 1 preliminary experiment. (E-G) Melting curves where analyzed subsequent to each qRTPCR to verify specific products have been built by displaying the change of the relative fluorescence units (RFU) with time. The melting curve of all samples (triplicates are shown in the same color) of one individual is presented here exemplary. (H) Organ distribution of rtOATP1d1 (squares) and the mean of two reference genes 18srRNA and Ef1α (circles) of all tested fish show equal organ distribution of reference genes.

Page 102: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL

Supplemental Figures

- 102 -

Supplemental Figure 2: Organ distribution of Oatp1d1 in (A) rainbow trout and (B) common carp using PCR with subsequent gel-electrophoresis (carried out with taq polymerase according to manufacteures protocol and gene specific primers for Oatp1d1). Elongation factor 1α (ef1α) was used as reference gene. Shown is a representative PCR from one individual. (Primers used: rainbow trout: Oatp1d1 5´-ATCGATGGCAGCTTTGAGAT-3´ and 5´-GGCAGTGTTCTCCTTTCTGG-3´; ef1α 5´-GTTCCGGCAAGAAACTTGAG-3´ and 5´-GACATCCTGTGGGAGGAGAA-3´; carp: Oatp1d1 5´-GGCCATGGGATCCTTCATTAC-3´ and 5´-GTCCACTGCTCCTATGTCC-3´; ef1α 5´-GTCGGTCGTGTTGAGACTGGTATCCT-3´ and 5´-ATCAGTTTGACAATGGCGGCATCT-3´.

Supplemental Figure 3: Expression-levels of Oatp1d1-mRNA in various organs of common carp (A), whitefish (B), rainbow trout (C) and Zebrafish (D) were determined using semi-quantitative real-time PCR. Raw threshold cycles obtained from the real time PCR of Oatp1d1 (●) and the respective housekeeping gene used for comparison (♦). For common carp (n=5) and zebrafish (n=3) 18s-rRNA has been used as reference gene, for whitefish (n=9) and rainbow trout (n=5) EF1α has been used as reference gene. All experiments were conducted in technical triplicates.

Page 103: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 105 -

Supplemental Figure 4: Transport of 6.1 nM [3H] E3S in HEK293 cells expressing Oatp1f1 (black bar) or Oatp1f2-1 (grey bar) was assessed two days after transfection and compared to the uptake into empty vector transfected HEK293 cells (white bar). Columns represent means ± SEM from three independent experiments (n=3) run in three technical replicates. No statistical difference was found between uptake of Oatp transfected cells and HEK293 cells containing the empty vector (one tailed Mann-Whitney). (A) 15 min uptake and (B) 30min uptake.

Page 104: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 106 -

Supplemental Figure 5: Alignment of Oatp1c1 amino acid sequence reported in chapter 4 as compared to the reference (ref) sequence (Acc.#: NP_001038462). Differing amino acids are highlighted.

Page 105: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 107 -

Supplemental Figure 6: Alignment of Oatp1d1 amino acid sequence reported in chapter 4 as compared to the reference (ref) sequence (Acc.#: NP_001082802). Differing amino acids are highlighted.

Page 106: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 108 -

Supplemental Figure 7: Alignment of Oatp1f1 amino acid sequence reported in chapter 4 as compared to the reference (ref) sequence (Acc.#: NP_000998082). Differing amino acids are highlighted.

Page 107: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 109 -

Supplemental Figure 8: Alignment of Oatp1f2 amino acid sequence reported in chapter 4 as compared to the reference (ref) sequence (Acc.#: NP_001121745). Differing amino acids are highlighted.

Page 108: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 110 -

Supplemental Figure 9: Alignment of Oatp1f3 amino acid sequence reported in chapter 4 as compared to the reference (ref) sequence (Acc.#: NP_001129156). Differing amino acids are highlighted.

Page 109: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 111 -

Supplemental Figure 10: Alignment of Oatp1f4 amino acid sequencereported in chapter 4 as compared to the reference (ref) sequence (Acc.#: NP_001074135). Differing amino acids are highlighted.

Page 110: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 112 -

Supplemental Figure 11: Alignment of Oatp4a1 amino acid sequencereported in chapter 4 as compared to the reference (ref) sequence (Acc.#: XP_696263). Differing amino acids are highlighted.

Page 111: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Figures

- 113 -

Supplemental Figure 12: Amino acid identities of the family Oatp1 in mammals and fish. Given in [%].Based on Clustal W alignment using ClustalW2 (http://www.ebi.ac.uk/Tools/msa/clustalw2/) and following multiple alignment options: Protein weight matrix: Gonnet; gap open: 10; gap extension: 0.2. The respective phylogenetic tree is shown in Figure 10. The letters A-F represent the respective Oatp subfamily. Species names in the left hand column are given as abbreviations in the first row (underlines letters).

Page 112: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Tables

- 115 -

SUPPLEMENTAL TABLES Supplemental Table 1: Accession numbers for sequences used in phylogenetic tree analysis

(Figure 3, chapter 2). species Oatp subtype Accession number

atlantic cod Oatp1d1 ENSGMOT00000018856

three-spined stickleback Oatp1d1 ENSGACT00000025261

fugu rubripes Oatp1c1-like* XP_003977582

rainbow trout Oatp1d1 KJ831065

zebrafish Oatp1d1 NP_001082802

atlantic cod Oatp1c1 ENSGMOT00000002810

three-spined stickleback Oatp1c1 ENSGACT00000011579

japanese medaka Oatp1c1-like* XP_004075747

nile tilapia Oatp1c1-like * XP_003443928

zebrafish Oatp1c1 NP_001038462

human OATP1C1 AAH22461

rhesus monkey OATP1C1* XP_002798534

cattle Oatp1c1 NP_001178438

house mouse OATP1C1 AAH78456

norway rat OATP1C1 NP_445893

chicken Oatp1c1 BAE72136

african coelacanth Oatp1c1-like* XP_005998930

little skate Oatp1d1 AAL66021

human OATP1A2 NP_066580

house mouse OATP1A5 AAH13594

zebrafish Oatp1f2 NP_001121745

zebrafish Oatp1f4 XP_005171769

zebrafish Oatp1f1 NP_998082

zebrafish Oatp1f3 XP_005174658

human OATP1B1 AAI14377

rhesus monkey OATP1B1* XP_001097704

human OATP1B3 EAW96414

rhesus monkey OATP1B3 NP_001028113

cattle Oatp1b3 NP_991373

norway rat OATP1B2 EDM01545

chicken OATP1B3 AEQ33635

western clawed frog Oatp1b3 NP_001027493

Page 113: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Tables

- 116 -

Supplemental Table 2: rtOatp1d1 specific primer pairs used for PCR and real time PCR. (Figure 4, chapter 2).

Real time PCR PCR

rtOatp1d1

5´-CTATTGTCTGTACGGCCTGG-3´

5´ATTGTTGTTCCCATTGGCCT-3´

5´-ATCGATGGCAGCTTTGAGAT-3´

5´-GGCAGTGTTCTCCTTTCTGG-3´

ef1α

5´-GCTTCAACGTCAAGAACGTG-3´

5´-CAGGTGGTTCAGGATGATG-3´

5´-GTTCCGGCAAGAAACTTGAG-3´

5´-GACATCCTGTGGGAGGAGAA-3´

18-srRNA

5´TCGCGTTGATTAAGTCCCTG-3´

5´-AGTTTGATCGTCTTCTCGGC-3´

Supplemental Table 3: Accession numbers for sequences used in phylogenetic tree analysis. (Figure 9-Figure 11, chapter 3)* predicted sequences from automated computational analysis.# Sequence used in the alignment in Figure 11 of chapter 2, but not in the phylogenetic trees of that study. species Oatp subtype Accession number

three-spined stickleback Oatp1d1* ENSGACT00000025261

common carp Oatp1d1 This study

japanese puffer fish Oatp1c1-like* XP_003977582

rainbow trout Oatp1d1 KJ831065

zebrafish Oatp1d1 NP_001082802

whitefish Oatp1d1 This study

green spotted puffer fish # Oatp1d1* CAF89838

atlantic cod Oatp1c1* ENSGMOT00000002810

three-spined stickleback Oatp1c1* ENSGACT00000011579

japanese medaka Oatp1c1-like* XP_004075747

nile tilapia Oatp1c1-like * XP_003443928

zebrafish Oatp1c1 NP_001038462

human OATP1C1 NP_059131

rhesus monkey OATP1C1* XP_002798534

cattle Oatp1c1 NP_001178438

house mouse OATP1C1 NP_067446

norway rat OATP1C1 NP_445893

chicken Oatp1c1 NP_001034186

african coelacanth Oatp1c1-like* XP_005998930

little skate Oatp 1d1 AAL66021

human OATP1A2 NP_066580

house mouse OATP1A5 NP_001254636

zebrafish Oatp1f2 NP_001121745

zebrafish Oatp1f4 NP_001074135

zebrafish Oatp1f1 NP_998082

zebrafish Oatp1f3* XP_005174658

human OATP1B1 NP_006437

rhesus monkey OATP1B1* XP_001097704

human OATP1B3 NP_062818

Page 114: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Tables

- 117 -

species Oatp subtype Accession number

rhesus monkey OATP1B3 NP_001028113

cattle OATP1B3 NP_991373

norway rat OATP1B2 NP_113838

chicken OATP1B3* XP_416418

human OATP3A1 NP_037404

rhesus monkey OATP3A1* AFE65910

cattle OATP3A1 NP_001001134

house mouse OATP3A1 NP_076397

norway rat OATP3A1 NP_803434

chicken OATP3A1* XP_413876

african coelacanth Oatp3a1* XP_005994028

three-spined stickleback Oatp3a1* ENSGACT00000014252

japanese medaka Oatp3a1* XP_004069706

atlantic cod Oatp3a1* ENSGMOT00000014299

japanese puffer Oatp3a1-like* XP_003969467

nile tilapia Oatp3a1-like* XP_003437749

zebrafish Oatp3a1 NP_001038653

african coelacanth Oatp2b1* XP_006007134

japanese puffer Oatp2b1-like* XP_003975969

nile tilapia Oatp2b1* XP_003457731

three-spined stickleback Oatp2b1* ENSGACT00000002454

japanese medaka Oatp2b1* XP_004076797

zebrafish Oatp2b1 NP_001032767

human OATP2B1 NP_009187

rhesus monkey OATP2B1 NP_001244739

cattle OATP2B1 NP_777268

house mouse OATP2B1 NP_001239459

norway rat OATP2B1 NP_542964

three-spined stickleback Oatp2a1* ENSGACT00000004896

zebrafish Oatp2a1 NP_001083051

house mouse OATP2A1* NP_201571

japanese medaka Oatp5a1-like* XP_004078845

zebrafish Oatp5a1-like* XP_684701

nile tilapia Oatp5a1* XP_003443615

three-spined stickleback Oatp5a1* ENSGACT00000003719

african coelacanth Oatp5a1* XP_005995937

chicken OATP5A1* XP_418287

house mouse OATP5A1 NP_766429

norway rat OATP5A1 NP_001101368

human OATP5A1 NP_112220

rhesus monkey OATP5A1* XP_001099455

European sea bass Oatp5a1* CBN81921

human OATP6A1 NP_775759

Page 115: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Tables

- 118 -

species Oatp subtype Accession number

rhesus monkey OATP6A1* XP_001095830

cattle OATP6B1-like* DAA27143

norway rat OATP6B1* EDL91883

house mouse OATP6C1 NP_083218

norway rat OATP6C1 NP_775460

house mouse OATP6D1 NP_001157705

african coelacanth Oatp4c1* XP_005993522

house mouse OATP4C1 NP_766246

norway rat OATP4C1 NP_001002024

human OATP4C1 NP_851322

rhesus monkey OATP4C1* XP_001097443

cattle OATP4C1 NP_001179775

japanese medaka Oatp4a1-like* XP_004070642

three-spined stickleback Oatp4a1* ENSGACT00000005884

atlantic cod Oatp4a1* ENSGMOT00000007689

zebrafish Oatp4a1* XP_696263

human OATP4A1 NP_057438

rhesus monkey OATP4A1* XP_001087529

cattle OATP4A1 NP_001179656

house mouse OATP4A1 NP_683735

norway rat OATP4A1 NP_683735

Supplemental Table 4: rtOatp1d1 specific primer pairs used for semi quantitative real time PCR (Figure 12, chapter 3).

Real time PCR-Primer

Whitefish Ef1α

5´-TGCCCCTCCAGGATGTCTAC-3´

5´-GGACACGTTCTTGACGTTA-3´

Whitefish Oatp1d1

5´-TCCAATCGGCAGAAGAAGTT-3´

5´-TCACCAAATACCCCAACACA-3´

Carp 18-srRNA

5´-GAGTATGGTTGCAAAGCTGAAAC-3´

5´-AATCTGTCAATCCTTTCCGTGTCC-3´

Carp Oatp1d1

5´-TGCGACACCACATTTCTTCC-3´

5´-GTGCAAGACTCCCATTGGTT-3´

Page 116: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Tables

- 119 -

Supplemental Table 5: Annotated Oatp sequence used for primer design (chapter 4) and resulting gene specific primer pairs with containing restriction sites (marked in grey) used for PCR amplification and ligation into the vector backbone.

Supplemental Table 6: Accession numbers of the zebrafish Oatp subtype sequences cloned in chapter 4.

Protein name Accession ID Oatp1d1 KP294314 Oatp1f1 KP294315 Oatp1f2-1 KP294316 Oatp1f2-2 KP294317 Oatp1f3 KP294318 Oatp1f4-1 KP294319 Oatp1f4-2 KP294320 Oatp2b1 KP294321 Oatp4a1-1 KP294322 Oatp4a1-2 KP294323

Protein name

Accession ID Primer (5’-3’) Restriction sites

organ of cDNA origin

Oatp1c1 NP_001038462 GCCCGCTAGCACCATGCCAAGGAACTGCATG GCGCGAATTCTTACAGCTGTTGTTTTCTTC

NheI EcoRI

brain

Oatp1d1 NP_001082802 GCGGGAATTCACCATGAGTACGGAGAAGAAG CGCCAGGATATCCAGGACTTCAGATGGTGGTC

EcoRI EcoRV

brain

Oatp1f1 NP_998082 GCGGGCTAGCAATATGTTGGTCATTACACTGG GCGGGAATTCTTATTGTTCTGCACTTTTTC

NheI EcoRI

kidney

Oatp1f2 NP_001121745 GCGGGCTAGCATCATGGAAGGCAGCACAATTGG GGCGGAATTCTTATTTAACAGCGTCCTCGTG

NheI EcoRI

kidney

Oatp1f3 NP_001129156 GCGGGCTAGCGAAATGGGTAATATACTGGTC GCCGGAATTCCCACAGATCATTATTTTATAGTCTCC

NheI EcoRI

kidney

Oatp1f4 NP_001074135 GCCGGCTAGCCATCATGGAAGGCAGCGC GCGGGAATTCTCAGTGTCTTGAGACATTTG

NheI EcoRI

kidney

Oatp2a1 NP_001083051 GCCCGCTAGCGACATGGACATATACGCCAAAG CACCGCGATATCTTATACAAAAAACGAATACC

NheI EcoRV

gut

Oatp2b1 NP_001032767 CGCCGCTAGCAAAATGACAGCATTCGAACC GACCGCGATATCCTAAGGACACGTTGTATTGC

NheI EcoRV

gut

Oatp4a1 XP_696263 GCCCGCTAGCGACATGCCCCACCTGTTGAAC GCGCGAATTCGGGTCATAATTTAGCGTG

NheI EcoRI

brain

Page 117: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Supplemental Tables

- 120 -

Supplemental Table 7: Predicted number of transmembrane domains of annotated Zebrafish Oatp sequences (Acession number is given) and sequences cloned in the study presented in chapter 4(marked with *). Three different server for prediction of transmembrane domains were consulted. TMHMM (http://www.cbs.dtu.dk/services/TMHMM/), which uses the hidden Markov model. Number in brackets represent additional amount of transmembrane domains with a probability <50 %. TMPred (http://www.ch.embnet.org/software/TMPRED_form.html). TOPCONS (http://www.topcons.net/) uses different methods or prediction and gives the resulting consensus number of predicted transmembrane domains.

transporter Putative number of transmembrane domains TMHMM TMPred TOPCONS

Oatp1c1 NP_001038462 10 10 12 * 11 11 12

Oatp1d1 NP_001082802 10 (2) 12 12 * 10 (2) 12 12

Oatp1f1 NP_998082 9 10 10 * 9 (1) 10 10

Oatp1f2 NP_001121745 11 (1) 12 12 *variant 1f2-1 11 (1) 12 12 *variant 1f2-2 8 (2) 10 10

Oatp1f3 NP_001129156 9 (1) 10 10 * 8 (1) 10 10

Oatp1f4 NP_001074135 9 (3) 12 12 *variant Oatp1f4-1 9 (3) 12 12 *variant Oatp1f4-2 9 (3) 12 12

Oatp2b1 NP_001032767 12 12 12 * 12 12 12

Oatp4a1 XP_696263 11 (1) 12 12 **variant Oatp1f4-1 11 (1) 12 12 *variant Oatp1f4-2 11 (1) 12 12

Page 118: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 121 -

NOVEL SEQUENCES IDENTIFIED IN THE PRESENT STUDY

Rainbow trout Oatp1d1 (2118 bp)

>coding sequence ATGAGTGTTGAAACTAAGGCGCCCCGGGTGCCTTGCTGCTCCAAACTGAAGCTGTTCCTGATTTCCCTGTCCTTTGTGTATTTCGCAAAAGCCTTCGGGGGAAGCTACATGAAGAGTTCTGTCACCCAGATAGAGAGACGTTTCGACATCCCCAGCTCACTTATAGGAGTCATCGATGGCAGCTTTGAGATTGGTAACCTGCTGGTGATAGCCTTTGTGAGCTACTTTGGGGCGAAGCTCCACAGGCCCAGACTGATAGGGATTGGCTGTCTCATTATGGCTGCTGGGGCCTTCCTAACGGCTCTGCCACACTTCTTCCAGGGACAGTATGTATATGAAACAACCCTGTCACATATACCTGAGGAAAACTCGACAGAAAGTATTCTGCCTTGTCTGGCCAACACGAGTCATCTGGTCATGAATGATCTGTCTACAGCTGCGAGTAAAGCAGCGTGTGAGAAGGAGGCAGGCTCGTCCATGTGGATCTATGTATTCCTGGGTAACATGCTGCGAGGGATCGGAGAGACGCCTGTCATGCCTTTAGGAGTGTCCTATCTGGACGACTACGCCAGAAAGGAGAACACTGCCATCTACCTTGCGTGTTTACAGACTGTTGGCATCCTGGGGCCTGTATTTGGCTTCATGTTTGGCTCCTTCTGTGCCAAGATCTACGTGGACATTGGCTCTGTGGATTTAGAGAGCATCTCCATCAACCACAAGGACTCTCGCTGGGTGGGGGCCTGGTGGTTAGGCTTCCTGGTTTCTGGCTTCCTGATGCTCCTGGCTGGGATTCCATTTTGGTTCCTCCCCAATACTCTGGCTAAGCAGTGCAAGAGCCCCAGCAAAAAGAGACAGAAGGAGGACATACAGGCAAAGAGCTCCTCTGACACCCCAGAGGTGGAGCCGGAGCAGGGCAGTTTCCTACCGGAGGACAACACTGAGGAGGATGCACCACAGCCTGTCACCATGGCTGCCATGGCTAAAGATTTTGCGCCGTCCCTGAAAAGGCTGTTCAGCAACAAAGTCTACCTGCTGATCATCCTGACGTCCGTGGTGGCATTCAATGGCTTCATAGGGATGATCACCTTCAAACCCAAGTTCATGGAGCAGGTCTACGGGCAGTCTCCCTCCAAGGCAATCTTCTTGATAGGTTCGATGAACCTGCCTGCGGTGGCAGTTGGGGTCATCGTGGGAGGCTTGGTGATAAAGCGGTTCAAGCTGAGCGTTCTGGGAGCCGCCCGACTCTCCATCGTCTCCTCCTTCCTCTCCTTCTGCCTCCTGCTCATCCAGTACTTCCTACAGTGTGACAACTCAGAGGTGGCTGGCCTCACCATGACCTATCAAGGGGCTACAGAGGTGTCTTACCAGCAGGAGAGCTTGCTATCTCAGTGTAACATGGGCTGCTCCTGTTCCCTGAAACACTGGGACCCTGTGTGTGCCAGAAATGGCATCACCTATGCCTCCCCCTGTCTTGCTGGCTGCCAGACCTCCACGGGCGTAGGCAAGGAGATGGAGTTCCATAACTGCACATGTATGCAGGAGCTGAAGGGGGGGCTGATGGCTCCACCAACCAACATGTCCGCCATCTTGGGACAGTGCCCCAGGAAGAGTGATTGTGACAGGATGTTTAAGTTCTACATGGCTGTCACTGTCCTGGGGGCCTTCGTCTCTGCCTGCGGGGGCACATCTGGATATATCGTCTTGCTCAGATCTATACACCCAGATTTAAAATCACTGGCTCTTGGGATGCATACATTGATAGGCAGAACTCTGGGTGGAATCCCTCCCCCTATCTACTTTGGAGCTTTGATTGACAGGACGTGTCTGAAGTGGGGATTGAAGCGATGTGGGGGTCAAGGAGCATGCAGATTCTACGACTCACATGCATTCAGAATAACGTTCCTGGGTCTGGTCTATTGTCTGTACGGCCTGGCCAACCTGCTGTGGGGGGTGCTGTACCTGCAACTGTCCAAGCGGCAGAAGAAGTTGGAACTGCGAAGCCAGGCCAAGGCTACAGCCCTGGAGGCCAATGGGAACAACAATCCCAATGGACATGCCATGGTCAGCATCTTTAAGAACAAAGATGACCGGGACAGGGACAATGAGAGCACCATCTGA

>translation

MSVETKAPRVPCCSKLKLFLISLSFVYFAKAFGGSYMKSSVTQIERRFDIPSSLIGVIDGSFEIGNLLVIAFVSYFGAKLHRPRLIGIGCLIMAAGAFLTALPHFFQGQYVYETTLSHIPEENSTESILPCLANTSHLVMNDLSTAASKAACEKEAGSSMWIYVFLGNMLRGIGETPVMPLGVSYLDDYARKENTAIYLACLQTVGILGPVFGFMFGSFCAKIYVDIGSVDLESISINHKDSRWVGAWWLGFLVSGFLMLLAGIPFWFLPNTLAKQCKSPSKKRQKEDIQAKSSSDTPEVEPEQGSFLPEDNTEEDAPQPVTMAAMAKDFAPSLKRLFSNKVYLLIILTSVVAFNGFIGMITFKPKFMEQVYGQSPSKAIFLIGSMNLPAVAVGVIVGGLVIKRFKLSVLGAARLSIVSSFLSFCLLLIQYFLQCDNSEVAGLTMTYQGATEVSYQQESLLSQCNMGCSCSLKHWDPVCARNGITYASPCLAGCQTSTGVGKEMEFHNCTCMQELKGGLMAPPTNMSAILGQCPRKSDCDRMFKFYMAVTVLGAFVSACGGTSGYIVLLRSIHPDLKSLALGMHTLIGRTLGGIPPPIYFGALIDRTCLKWGLKRCGGQGACRFYDSHAFRITFLGLVYCLYGLANLLWGVLYLQLSKRQKKLELRSQAKATALEANGNNNPNGHAMVSIFKNKDDRDRDNESTI*

Page 119: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 122 -

Common carp Oatp1d1 (2064 bp)

>coding sequence

ATGAGTGTGGAAAAGAAGACGCCACAAGAGCCATGCTGCTCCAAATTAAAGATGTTTTTGGCTGCCATGTGTTTTGTGTACTTTGCCAAAGCATTTCAGGGCAGCTACATGAAGAGCTCTATCACACAGATTGAGAGACGTTTTGACATTCCAAGCTCACTGATTGGATTAATTGATGGAAGTTTTGAAGTTGGTAATCTTTTGGTAATAGCTTTTGTGAGTTACTTTGGTGCAAAGCTTCATAGACCACGACTCATAGGAGCTGGATGTTTAATTATGGCCATGGGATCCTTCATTACTGCGGCACCACATTTCTTCCAAGGACTGTACAAATACGAGACAACAGTGTCACATTTCTCTGCCTCTAATGGTACAGAGAGTATTCTACCATGCTTAACCAATGGGAGTCTTGCACAAGATGAACTTCCCACAGTGGAAATTCAAGCAGAATGTGAGAAAGCAGCCAGTTCCTCCTTGTGGCTCTATGTGTTTCTTGGGAACATGCTTCGTGGGGTAGGAGAGACACCTGTCATGCCCCTTGGGTTGTCTTATCTAGATGATTTTTCAAGGGAGGAAAACACTGCTCTTTACATGGCTCTCATACAGACTGTGGGAATAATGGGGCCTATGTTTGGATTCATGTTGGGATCTTTCTGTGCCAAGCTCTACGTGGACATAGGAGCAGTGGACTTAGATACCATTTCCATCAACCACAAAGATTCACGTTGGGTTGGTGCCTGGTGGCTTGGCTTTTTGTTAACTGGTGGAGTGATGTTACTGGCTGGAATTCCATTTTGGTTCCTTCCCAAGTCTCTTCCTAAGCAGGGTGAGACTGAATCTGAGAAAAAATTGAATGAGGGCGAGCAGGACTGTTTCATTCCCGACAACAACAAGCACAGTAATGCTCCTGACAAACCAGCTCCGGTCACCATGGCAGCTCTGGCTAAAGATTTTCTGCCATCACTGAAAAAACTCTTCAGCAACCAGATTTACCTGCTCCTTGTATGCACTGCATTTGTGCAAGTCAATGGTTTTATAGGAATGATCACATTTAAGCCAAAGTTTATGGAGCAAATTTATGGCCAGTCAGCATCAAAGGCCATTTTTTTTATAGGCATTATGAATCTACCAGCTGTGGCTTTGGGAATTGTCACCGGTGGGTTTATAATGAAGAAGTTCAAACTGAATGTTCTCGGAGCAACCAAGATCTGCATTGGAACATCTCTTCTGGCTTTCTTTACGATGCTTATTCAGTATTTCCTGCAGTGTGACAACACACAAGTGCCAGGACTTACATTGTCATATCAAGGGGTTCCACAGGTGTCCTATCAGCAGAATACGCTAATTTCTCCATGTAACATGGGTTGCTCCTGCTCTCTGAAACATTGGGATCCCATATGTGCCAGCAATGGCTTGACCTATGCCTCCCCCTGCCTTGCTGGCTGCCAAAGCTCCACTGGTGATGGCAAGAATATGGTATTCCACAACTGCACCTGTGTTGGAGATTCGCCTTTCCCATATGCAAACATGTCAGCAGTGCTGGGCCAGTGCCCTCGCAAGAGTGATTGTGACTACATGTTTAAGGTCTACATGGCGGTGACAGTCATTGGTGCCTTTCTCTCGGCTTGTGGGGCCACACCAGGTTACATTATTCTACTCAGATCCATAAATCCAGAACTAAAATCTTTGGCTCTTGGTATTTATACTTTGATTGTTCGGACTCTGGGTGGCATTCCCCCTCCAGTTTATTTTGGAGCCTTTATTGACAGGACCTGTCTGAAATGGGGCACAAAACAATGTGGAGGGAGAGGAGCGTGTCGTATCTATGATTCTGGTGCCTTCAGAAACGCTTTCTTGGGTCTTATTTATGGCCTATATTCCTTATCCTACATATTGTGGGGAGTCGTGTACATTAGACTGTCTCATCGTGAAAAGAAGCTTGCACTGAAGAACCAGTTAAAAACTCCAGAACAAGATGGCAATGGTGTTTCTGCAGGGAATGGCAATGCATCTTCAGCTATTGTAAAATGCGGTGAAAACGCAGACCAGGAGACCACTATCTGA

>translation

MSVEKKTPQEPCCSKLKMFLAAMCFVYFAKAFQGSYMKSSITQIERRFDIPSSLIGLIDGSFEVGNLLVIAFVSYFGAKLHRPRLIGAGCLIMAMGSFITAAPHFFQGLYKYETTVSHFSASNGTESILPCLTNGSLAQDELPTVEIQAECEKAASSSLWLYVFLGNMLRGVGETPVMPLGLSYLDDFSREENTALYMALIQTVGIMGPMFGFMLGSFCAKLYVDIGAVDLDTISINHKDSRWVGAWWLGFLLTGGVMLLAGIPFWFLPKSLPKQGETESEKKLNEGEQDCFIPDNNKHSNAPDKPAPVTMAALAKDFLPSLKKLFSNQIYLLLVCTAFVQVNGFIGMITFKPKFMEQIYGQSASKAIFFIGIMNLPAVALGIVTGGFIMKKFKLNVLGATKICIGTSLLAFFTMLIQYFLQCDNTQVPGLTLSYQGVPQVSYQQNTLISPCNMGCSCSLKHWDPICASNGLTYASPCLAGCQSSTGDGKNMVFHNCTCVGDSPFPYANMSAVLGQCPRKSDCDYMFKVYMAVTVIGAFLSACGATPGYIILLRSINPELKSLALGIYTLIVRTLGGIPPPVYFGAFIDRTCLKWGTKQCGGRGACRIYDSGAFRNAFLGLIYGLYSLSYILWGVVYIRLSHREKKLALKNQLKTPEQDGNGVSAGNGNASSAIVKCGENADQETTI*

Page 120: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 123 -

Whitefish Oatp1d1 (2118 bp)

>coding sequence ATGAGTGTTGAAACTAAGGCGCCCCGGGTGCCTTGCTGCTCCAAACTGAAGCTGTTCCTGATTTCCCTGTCCTTTGTGTATTTCGCAAAAGCCTTCGGGGGAAGCTACATGAAGAGTTCTGTCACCCAGATAGAGAGACGTTTCGACATCCCCAGCTCACTTATAGGAGTCATCGATGGCAGCTTTGAGATGGGTAACCTGCTGGTGATAGCATTTGTGAGTTACTTTGGGGCGAAGCTCCACAGGCCCAAACTGATAGGGATTGGCTGTCTCATCATGGCTGCTGGGGCCTTCCTCACCGCTCTGCCACACTTCTTCCAGGGAAAGTACGTATATGAAACAAACCTGTCACATATACTTGAGGAAAACACTACAGAGAGCATTCTGCCTTGTCTGGCCAACACAAGTCATCTGGTCATAAATGATCAGGCTACAGCTGCGAGTAAAGCAGCATGTGAGAAGGAGGCAGGCTCGTCCATGTGGATCTATGTATTCCTGGGTAACATGCTGCGTGGGATTGGAGAGACGCCTGTCATGCCTTTAGGAGTATCCTATCTGGACGACTACGCCCGAAAGGAGAACACTGCCATCTACCTTGCGTGTTTACAGACTGTTGGCATCCTGGGGCCCGTATTTGGCTTCATGTTTGGCTCCTTCTGTGCCAAGATCTATGTGGACATTGGCTCTGTGGATTTAGAGAGCATCTCCATCAACCACAAGGACTCTCGCTGGGTGGGGGCCTGGTGGTTAGGCTTCCTGGTATCTGGCTTCCTGATGCTCCTGGCTGGGATTCCATTTTGGTTCCTCCCCAATACTCTGCCTAAACAGGGCAAGAGCCCCAGCAAAAAGAAACAGAAGGATGACATACAGGCCAAGAGCTCATCTGACACCCCAGAGGAGGAGCAGGAGCAGGGCAGTTTCCTACCAGAGGACAATACTGAGGAGGATGCACCAAAGCCTGTCACCATGGCTGCCATGGCTAAAGATTTTGGGCCGTCCCTGAAAAGACTGTTCAGCAACAAAGTCTACCTGCTGATCATCCTGACATCTGTGGTGGCATTTAACGGCTTCATAGGGATGATCACCTTCAAACCTAAGTTCATGGAGCAGGTCTACGGGCAGTCTCCCTCCAGGGCAATCTTCTTGATTGGTTCGATGAACCTGCCTGCGGTGGCAATTGGGATCATCGTGGGAGGCTATGTGATGAAGCGGTTCAAGCTGAGTATTCTGGGAGCAGCCCGACTCTCCATCATCTCCTCCTTCCTCTCCTTCTGCCTCCTGCTCATCCAGTACTTCCTACAGTGTGACAACTCAGAGGTGGCTGGCCTCACCATGACCTATAAAGGGGCTCCAGAGGTGTCTTACCAGCCGGAGAGCTTGCTGTCTCAGTGTAACATGGGCTGCTCCTGTTCCCTGAAACACTGGGACCCTGTGTGTGCCAGCAATGGCATGACCTACGCCACCCCCTGTCTAGCTGGCTGCCAGACCTCCACGGGCGTAGGCAAGGAGATGGAGTTCCATAACTGCACATGTATGGGGGAGCTGAAGGGGCAAATGTTGGCTCTTCCAACCAACATGTCAGCCATTTTGGGCCAGTGCCCCAGGAAGAGTGACTGTGACAGGATGTTTAAGTTCTACATGGCTGTCTCTGTCCTGGGGGCCTTCGTCTCTGCCTGTGGGGGCACATCTGGATATATCGTCCTGCTCAGATCTATACACCCCGATTTAAAATCACTGGCTCTTGGCATGCAGACATTGATAGTCAGAACTCTGGGTGGAATCCCTCCCCCTATATACTTCGGAGCGTTGATTGACAGGACATGTCTGAAGTGGGGTTTAAAGCGATGTGGGGGTCAAGGAGCATGCAGACTCTACGACTCACATGCATTCAGAATAACGTTCCTGGGTCTGATCTATGGTCTATACGGCCTGGCCTACCTGCTGTGGGGGGTGCTGTACCGGCAGCTGTCCAATCGGCAGAAGAAGTTGGTACTGCGAAGCCAGGCCAAGGCTACAGCCCTGGAGGCCAACGGGAACAACAATCCCAATGGACATGCCATGGTCAGCATCTTTAAGAACAAAGATGATCTGGACAGGGACAGTGAGAGCACTATCTGA

>translation

MSVETKAPRVPCCSKLKLFLISLSFVYFAKAFGGSYMKSSVTQIERRFDIPSSLIGVIDGSFEMGNLLVIAFVSYFGAKLHRPKLIGIGCLIMAAGAFLTALPHFFQGKYVYETNLSHILEENTTESILPCLANTSHLVINDQATAASKAACEKEAGSSMWIYVFLGNMLRGIGETPVMPLGVSYLDDYARKENTAIYLACLQTVGILGPVFGFMFGSFCAKIYVDIGSVDLESISINHKDSRWVGAWWLGFLVSGFLMLLAGIPFWFLPNTLPKQGKSPSKKKQKDDIQAKSSSDTPEEEQEQGSFLPEDNTEEDAPKPVTMAAMAKDFGPSLKRLFSNKVYLLIILTSVVAFNGFIGMITFKPKFMEQVYGQSPSRAIFLIGSMNLPAVAIGIIVGGYVMKRFKLSILGAARLSIISSFLSFCLLLIQYFLQCDNSEVAGLTMTYKGAPEVSYQPESLLSQCNMGCSCSLKHWDPVCASNGMTYATPCLAGCQTSTGVGKEMEFHNCTCMGELKGQMLALPTNMSAILGQCPRKSDCDRMFKFYMAVSVLGAFVSACGGTSGYIVLLRSIHPDLKSLALGMQTLIVRTLGGIPPPIYFGALIDRTCLKWGLKRCGGQGACRLYDSHAFRITFLGLIYGLYGLAYLLWGVLYRQLSNRQKKLVLRSQAKATALEANGNNNPNGHAMVSIFKNKDDLDRDSESTI*

Page 121: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 124 -

Zebrafish Oatp1d1 (2070 bp)

>coding sequence ATGAGTACGGAGAAGAAGAAGGAGCCATGCTGCAGCAAGTTAAAGATGTTTTTGGCTGCCATGTGCTTTGTGTTCTTTGCCAAAGCCTTTCAGGGCAGCTACATGAAGAGCTCCGTCACGCAGATCGAGAGGCGTTTTGACGTCCCCAGCTCGCTAATCGGATTCATCGATGGAAGTTTTGAAATCGGTAATCTGTTTGTAATAGCTTTCGTGAGTTACTTTGGTGCTAAGCTTCACAGGCCGCGGCTGATAGCAGCTGGATGTTTAGTCATGTCCGCAGGATCCTTCATAACAGCGATGCCGCATTTCTTTCAAGGACAGTACAAATACGAGTCAACCATTTCTCATTTCTCGGCCTCGGTAAACGGTACAGAGAATGTTCTGCCGTGCCTTACAAATGCGAGTCTTGCACAAGATAGTGAAATTCCTACTGTGGAGTCTCAAGCAGAGTGTGAAAAAGCCTCCAGTTCTTCCTTGTGGCTGTTTGTTTTTCTGGGAAACATGCTTCGTGGGATTGGAGAGACGCCTGTCATGCCTCTCGGCTTGTCTTACCTGGATGATTTTTCAAGAGAGGAAAACACTGCTTTTTACCTGGCTCTCATACAGACGGTGGGAATCATGGGGCCGATGTTTGGATTCATGCTGGGCTCGTTCTGTGCCAAGCTGTACGTGGACATTGGAACGGTGGACTTAGACAGCATCACCATCAACTATAAGGACTCGCGTTGGGTCGGCGCCTGGTGGCTGGGCTTCTTGGTAACCGGTGGAGTGATGCTGCTGGCCGGAATCCCGTTCTGGTTTCTTCCCAAGTCTCTGACCAGACAGGGAGAGCCTGAAAGTGAGAAGAAACCAGGTGCTCCAGAAGGAGGAGAGCAGGAGCGCTTCATTCCTGACAACAACAAACACAATCCTCCTGCCAGCAAACCGGCGCCGGTCACCATGTCAGCCCTGGCTAAAGATTTTCTGCCGTCACTGAAGAAACTCTTCAGCAACACGATCTATGTGCTCCTGGTTTGCACCGGTTTGATACAAGTCAGTGGTTTCATAGGGATGATCACGTTCAAGCCGAAGTTCATGGAGCAAGTTTACGGCCAATCGGCATCAAGAGCAATATTCTTGATAGGCATCATGAATCTCCCAGCCGTGGCTCTGGGAATCGTCACTGGAGGGTTTATAATGAAGAGATTTAAGGTGAATGTTCTCGGAGCGGCCAAGATCTGCATCGTGGCGTCGGTTCTGGCTTTCTGTTCAATGCTCATCCAGTATTTCCTACAGTGTGACAACTCACAAGTGGCGGGACTTACAGTCACGTACCAAGGGGCTCCAGAAGTGTCGTACCAGACGGAAACGCTCATCTCTCAGTGCAACATTGGCTGCTCTTGCTCTCTGAAGCACTGGGATCCCATTTGTGCCAGCAATGGTGTGACCTACACCTCCCCCTGCCTCGCTGGCTGCCAGACCTCTACTGGGATTGGCAAGGAGATGGTTTTCCACAACTGCAGCTGTATTGGAGAAGCTCTTCTTCCGTACACGAACATGTCGGCAGTGTTGGGTCAGTGTCCGCGCAAGAGCGACTGCGACTTCATGTTTAAGATCTACATGGCGGTGACGGTTATCGGCGCCTTTTTCTCAGCTGTCGGAGCCACACCGGGTTACATTATTCTGCTCAGATCCATCACACCAGAACTGAAATCTTTGGCTCTCGGTATGCACACTTTGATTGTTCGCACTCTGGGTGGGATTCCCCCTCCGATATATTTCGGAGCCTTGATTGACAAGACCTGTCTGAAATGGGGTTTAAAGCAATGCGGCGGCAGAGGAGCGTGTCGAATCTACGACTCCGGAGCCTTCAGGAATGCTTTCTTGGGGTTAATTTACGCCCTGTACAGCTCATCCTACCTGCTGTTTGGACTTCTCTACAACAGACTGTCTCATCGAGAAAAAAAGCAAGCGCTAAAGGACCAGTTAAAAGCTCCAGAGCAGGACAACTGTGGTGTTTCAACCACCAATGGAAATGCGTCTTCGGCGATCGTAAAATGCGAAAACCCAGACCAGGAGACCACCATCTGA

>translation

MSTEKKKEPCCSKLKMFLAAMCFVFFAKAFQGSYMKSSVTQIERRFDVPSSLIGFIDGSFEIGNLFVIAFVSYFGAKLHRPRLIAAGCLVMSAGSFITAMPHFFQGQYKYESTISHFSASVNGTENVLPCLTNASLAQDSEIPTVESQAECEKASSSSLWLFVFLGNMLRGIGETPVMPLGLSYLDDFSREENTAFYLALIQTVGIMGPMFGFMLGSFCAKLYVDIGTVDLDSITINYKDSRWVGAWWLGFLVTGGVMLLAGIPFWFLPKSLTRQGEPESEKKPGAPEGGEQERFIPDNNKHNPPASKPAPVTMSALAKDFLPSLKKLFSNTIYVLLVCTGLIQVSGFIGMITFKPKFMEQVYGQSASRAIFLIGIMNLPAVALGIVTGGFIMKRFKVNVLGAAKICIVASVLAFCSMLIQYFLQCDNSQVAGLTVTYQGAPEVSYQTETLISQCNIGCSCSLKHWDPICASNGVTYTSPCLAGCQTSTGIGKEMVFHNCSCIGEALLPYTNMSAVLGQCPRKSDCDFMFKIYMAVTVIGAFFSAVGATPGYIILLRSITPELKSLALGMHTLIVRTLGGIPPPIYFGALIDKTCLKWGLKQCGGRGACRIYDSGAFRNAFLGLIYALYSSSYLLFGLLYNRLSHREKKQALKDQLKAPEQDNCGVSTTNGNASSAIVKCENPDQETTI*

Page 122: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 125 -

Zebrafish Oatp1c1 (2075 bp)

>coding sequence ATGCCAAGGAACTGCATGGAAGAGTCATCATCTCCCAAAGAAAACGAGATCTTTCTGTGGAGGATGGATCTGCCCAACAAGGAAAGACGGTCACCGGAAAACAGCCCAAGCTCCAGTCGGCGCTGTTGCTCAAGCTTCAAGATGTTCCTCGTGGCCTTGTCCTTCGCCTATTTCTCCAAGGCTTTATCCGGGAGCTATATGAAAAGCACAATAACTCAGCTGGAAAGACGCTTCGACATCCCCAGTTACTTAATAGGTGTCATTGACGGCAGCTTTGAAATAGGCAACTTGTTAGTGATCGCCTTTGTGAGTTACTTTGGTGCCAAACTTCACCGTCCGAAGATCATCGCCATCGGCTGCCTGTTGATGTCACTCGGGACGTTTCTCATCGCTTTGCCCCATTTCATAATCGGCCGATATAAGTTTGAGACGTCGATCCGCTTGTCTGTGAACTCCAGCAGCAGCCTCTCTCCGTGTTCACTGAGCTCTGCTGAACCCAGCACATCTTCAGTGCCCAGCACAGGCTGTGAACACGAGTCTCGGTTGTCGATGTGGATTTATGTGTTTCTGGGGAATATCCTGCGTGGGATTGGAGAGACTCCGGTGCAGCCGCTGGGAATCTCCTACATCGATGATCACGCTCTGGAGGAGAACGCAGCTTTCTACATCGGATGTGTCCAGACTATCTCCGTCATCGGGCCGGTTTTCGGGTATCTGCTGGGCTCCGTCTGTGCCAAGATATACGTGGACATTGGGTTTGTCAACATGGAGAGCATCAGCATCACTCCAGGAGACGCGCGCTGGGTTGGCGCCTGGTGGCTAGGCTACCTGATTGCCGGATTCATTACCCTGCTTTCAGCCGTGCCCTTCTGGTTCCTGCCCAAGTCTTTGCCACTGCCAGAGAGACGACTGAGCAAATACTCTCCGGAGCGTAACAGCTTCATCAAAGACTCGGCTCTGCTGGAGCACAAATACCAGGCAGACGAGCCGGCTAACTTCCTGGAAATGGCCAAAGACTTTTTGCCCACACTGAGGTCTCTCCTGGGAAACCCGGTCTACTTTCTGTATCTGTGCGTGACCATCATCCAGTTCAACTCACTGATCAGCATGGTCACATATAAACCAAAATACATCGAGCAGCATTACGGACAGTCCGCATCCAAAACCAACTTCCTCATGGGAGTTATCAACATCCCAGCCGTAGCCTTGGGTATGTTTTCGGGCGGCGTGATCATGAAGAAGTTCAAGCTGAGCATCATGGGAGCAGCCAAGTTTGTGTTGGGAACGTCTCTGCTGGGCTATTTTCTGTCTCTCTTCTTCTTCAGCATGGGCTGTGAAAACGCCAGCGTGGCTGGAATCACCATGTCCTACAACGGGACGGAGAGTCTGTTTATGGCGGACGGCTCTCTGCGAGCGTCCTGCAACGCAGCCTGTCATTGTCCTGAAAAAAACTGGGACCCGGTGTGTGGAGAAGACGGCCTCACCTACGTGTCCCCGTGCCTGGCTGGATGCCGGACGTCCCACGGCTCAGGAATGGACACGGTGTTTGAGCAGTGCAGTTGTATAGGAGTGGGGAATCAGACAGCCACCGCAGGCCAGTGCAACGACAGAGAAAGCTGCCATCGCATTTTCCCGTACTTCCTGGCTCTGTCCGTCATCACTTCCTTCATTATATCACTAGGAGGAACGCCAGGATACATGCTGCTCATCAGATGCATCAAACCTCAGCTGAAGTCTTTAGCGTTGGGTTTTCACACTCTGACCACACGTACTCTGGCTGGGATTCCGGCTCCGATATATTTCGGGGCCATCATCGACACCACCTGTCTGAAATGGGGACACAAGAAGTGTGGTGGAAGAGGAGCCTGTCGAATCTACAACACCGCGGCCTACAGGATAGCGTATCTGGGCCTGACGTTGGGTTTGCGCACCGTTTCCTTCTTCCTGTGTATCCTGGGGTTGTTGGTTTTGCGACAGCACGTTCAGAAGCAGGAGCGCAGCACAATGGTGGACGGAGGATCAGCGGCAGCCGGAGAACTGCAGGCGCTCAGAAAAGAAGAAAACAACAGCTGTAA

>translation

MPRNCMEESSSPKENEIFLWRMDLPNKERRSPENSPSSSRRCCSSFKMFLVALSFAYFSKALSGSYMKSTITQLERRFDIPSYLIGVIDGSFEIGNLLVIAFVSYFGAKLHRPKIIAIGCLLMSLGTFLIALPHFIIGRYKFETSIRLSVNSSSSLSPCSLSSAEPSTSSVPSTGCEHESRLSMWIYVFLGNILRGIGETPVQPLGISYIDDHALEENAAFYIGCVQTISVIGPVFGYLLGSVCAKIYVDIGFVNMESISITPGDARWVGAWWLGYLIAGFITLLSAVPFWFLPKSLPLPERRLSKYSPERNSFIKDSALLEHKYQADEPANFLEMAKDFLPTLRSLLGNPVYFLYLCVTIIQFNSLISMVTYKPKYIEQHYGQSASKTNFLMGVINIPAVALGMFSGGVIMKKFKLSIMGAAKFVLGTSLLGYFLSLFFFSMGCENASVAGITMSYNGTESLFMADGSLRASCNAACHCPEKNWDPVCGEDGLTYVSPCLAGCRTSHGSGMDTVFEQCSCIGVGNQTATAGQCNDRESCHRIFPYFLALSVITSFIISLGGTPGYMLLIRCIKPQLKSLALGFHTLTTRTLAGIPAPIYFGAIIDTTCLKWGHKKCGGRGACRIYNTAAYRIAYLGLTLGLRTVSFFLCILGLLVLRQHVQKQERSTMVDGGSAAAGELQALRKEENNSC*

Page 123: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 126 -

Zebrafish Oatp1f1 (1686 bp)

>coding sequence ATGTTGGTCATTACACTGGTAAGCTATGTTGGCGCTAAGTTTCACAGGCCCAAGATCATTGGAGCAGGGGTTCTGCTGATGGGTATAGGAACACTGCTAATGGCTTCACCTCATTTCATTATGGGCCGGTACAAGTATGGTACGGCTGCCACCCACACTAATGATGCTGGTAATTTCACTGTGATCTCTACATGCTCATCTGACTCTCAGGAAACTCTTCAACAGCCTTACTCTGAGTGCCAAATAATGGAAGCGGAAAGCTCACTTTTGTGGGTGGTAGTGCTTTTGGGAAACGCTATGCGTGGGATTGGGGAAGCCAGCATTGTCCCTCTGGGGATGTCATTCACAGATGATTATGCACGTCCAGAAAACTCTGCCTTTTACATTGGTTGTCTGAACACACTCAGAGGGATTGGGCCAATTTTTGGTTTCATTCTTGGATCTCTATGTGCTAATCTTTATGTGGACTTTGACTTAGTGAATCAAGAAAGCGTGACCATCACACCGCAGGACTCTCGCTGGGTTGGGGCTTGGTGGTTGGGTTTTGCGGTGTCTGGCCTGTTGACTCTCCTCGCTGCTTTTCCTTTCTGGTTCTTGCCAAGAGCTCTACCTGAAAACTCCCAAACCTCCGAGCTAGACCACCCCCAACAACAGCACAAAACCACACCCAGCCTTACAGAGATAGCCAAAGATTTTGCACCAACTTTTAAGCGTCTGCTGACCAATAAGATCTACATTCTGTACCTGGCGTACAGTATAGTGGCATTCAACGACTTCGCCATAGTTGCAACGTACACGCCAAAGTATCTGGAACAGCAGTTTGGGCAAAGTGCATCCAAAGCCAACTTTCTGATAGGAGTGACATGTATTCCGGCAGTGGCTGTGGGTATTTTCCTGAGTGGGTTGATGATGAAGAGGTTTAAATGGGGTCTGCTGGCATCCGCAAGAGTGAATTTGTTCACCGGTGTCACCATGTTGCTTTTGGCTGTGCCTTTCTTCGCTCTCAGCTGTGAGAATCTAGATGTCGCAGGGGTTACAGTGCCCTATCAAGGATCTGCTGAAGTCCAAGGTGTAGTAAGCGATGTACTCCCTTCTTGCAATGCTGACTGTGGGTGTCCAGAACTCCAATGGGATCCAGTGTGTGGAGAGAATGGACTGACCTACATCTCCCCTTGCCATGCGGGCTGCAATTCCACCCAGGGTGCAGGATGGAATAAGACATTCCATGACTGCAGGTGTATTCAGAGCTGGGGACTGAGCGTTGGCAACTCCTCTGCAGTTCTTGGACAGTGTTCACGAGATCCCAACTGCAACCAAATGATCTACGTTTTTCTGGGAGTGCAGTCTTTGGTCTTGTTTGTGTACAGTCTTGGCGCTGTTCCCTTTCTCACTTTTTCTATGAGGATTGTGGATCCTGAGCTGAAGGCTCTCTCTGTGGGAGTGTTACTGTTATCTATAAGAGTTCTAGGTGGTATTCCTGCACCCATTTACTTTAGTGGTCTCATTGACACCACCTGTCTGAAATGGGGCCAGAGGAAATCCTGTGGAAGGGGTGCCTGCAGAATTTACGACAATGAGACCTTCAGGTTCCTCTTTCAGGGAATGACCATCGGTCTTCGAGTATTAGCCTGTTCCGTGCTCTGGATAGCCACTATAGAGATAAAGAGAAAAAGTGCAGAACAATAA

>translation

MLVITLVSYVGAKFHRPKIIGAGVLLMGIGTLLMASPHFIMGRYKYGTAATHTNDAGNFTVISTCSSDSQETLQQPYSECQIMEAESSLLWVVVLLGNAMRGIGEASIVPLGMSFTDDYARPENSAFYIGCLNTLRGIGPIFGFILGSLCANLYVDFDLVNQESVTITPQDSRWVGAWWLGFAVSGLLTLLAAFPFWFLPRALPENSQTSELDHPQQQHKTTPSLTEIAKDFAPTFKRLLTNKIYILYLAYSIVAFNDFAIVATYTPKYLEQQFGQSASKANFLIGVTCIPAVAVGIFLSGLMMKRFKWGLLASARVNLFTGVTMLLLAVPFFALSCENLDVAGVTVPYQGSAEVQGVVSDVLPSCNADCGCPELQWDPVCGENGLTYISPCHAGCNSTQGAGWNKTFHDCRCIQSWGLSVGNSSAVLGQCSRDPNCNQMIYVFLGVQSLVLFVYSLGAVPFLTFSMRIVDPELKALSVGVLLLSIRVLGGIPAPIYFSGLIDTTCLKWGQRKSCGRGACRIYDNETFRFLFQGMTIGLRVLACSVLWIATIEIKRKSAEQ*

Page 124: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 127 -

Zebrafish Oatp1f2-1 (1989 bp)

>coding sequence ATGGAAGGCAGCACAATTGGCATGGAACATTCAGACAGAAACCAGAAAACATCTCAAACTAAATGCTGCCCTCCAAGGCTGAAGATCTTTATTTGCGCACTTGCCTTCTGTTACTTCTCTAAATCCATGACGGCGTCTTACACCAAAAGCACCATCACTCAAATCGAGAGGCGGTTTGAGATCCCGAGCTCCACCGTTGGCATCATTGATGGGGGTTTTGAAATGGGTAATATGCTGGTCATTACCCTGGTGAGCTACATTGGTGCTAAGTTTCACAGGCCGAAGATCATTGGAGCAGGGGTTCTGCTGATGGGTATAGGAACACTGCTAATGGCTTCACCTCATTTCATTATGGGCCGGTACAAGTATGGTACAGCTGCCACCCACACTAATGATGCTGGTAATTTCACTGTGATCTCTACATGCTCATCCGACTCTCAGAAAACTCTTCAACAGCCTTTTTCTGGATGCCGAAAAGAGGAAGCTGAAAGCTCACCCTTGTGGGTTATAGTGGTTTTGGGAAACACTATGCTTGGGATTGGGGAAGCCAGTATTATCCCTCTGGGAATGTCATTCGTAGATGATTATGCACGGCCAGAAAACTCAGCTTTTTACATTGGTTGTCTGAACACAATTAAAGGGATTGGGCCAATTTTTGGTCTCGTGCTTGGATCTCTTTGTGCTAATCTTTATGTGGATATTGGCCTAGTGAATAAAGAAAGCGTGACCATCACACCACAGGACTCTCGCTGGGTTGGGGCTTGGTGGTTGGGTTATGTGGTGTCTGGCTTGTTGACTGTCCTCGCTGCTTTTCCTTTCTGGTTCTTGCCAAAAGCTCTGCCTGAAAACTCCCAAATCTCACTGCTAGACAACACCCCACAACAGCACAAAACCACACCCAGCCTTACAGAGATAGTCAAAGATTTTGCGCCAACTTTTAAGCGTCTGCTGACCAATAAGATCTACATTCTGTGCCTGGCGTACAGTATAGTGGCATTCAACGACTTCGCCATACTTGTAACGTACACGCCAAAGTATCTGGAACAGCAGTTTGGGCAAAGCGCATCCAAAGCCAACTTTCTGATAGGAGTGACATGTGTTCCGGCAGTAGCGCTGGGTGTTTTCCTGAGTGGGTTGATGATGAAGAGGTTTAAATGGGGTCTGCTGGCATCCGCAAGAGTGAATTTGTTGACCATAGTCGCCACAGTGTTTTTAACCGTACCTTTCTTCGCTCTCAGCTGTGAAAATCTAGATATCGCTGGGGTTACAGTGCCCTATCAAGGATCTACTGAAGTTCAAGGTGTAACCAGTGATGTACTCCCTTCTTGCAATGCTGACTGTGGGTGTCCAGACCTCCATTGGGATCCAGTGTGTGGAGAGAATGGAGTGACCTACATCTCCCCTTGCCATGCGGGCTGCAGTTCCACCCAGGGTGCAGGACGGAATAAGACATTCCATGACTGCGGGTGTATTCAGAGCTGGGGACTGAGCGTTGGCAACTCCTCTGCAGTTCTTGGACAGTGTTCACAAAATCCCAACTGCAACAGAATGATGTACCTTTATCTGGGATTGCAGTCTTTGGCCCTCTTTGTGTACAGTCTTGGTGCTGTTCCCCTTTTCACCATGTCTCTGAGGATTGTGGATCCCGAGCTGAAGTCTCTCTCTGTGGGAGTGTTACTGTTATCTATAAGAGTTTTGGGTGGTATTCCTGCTCCCATTTACTTTGGTGCTCTTGTTGACTCCAGCTGTCTGAAATGGGGCCAGAGGAAATCTTGTGGAAGGGGTGCCTGCAGAATTTATGACATTGAGACGTTCAGGTTCCTCTTTCAAGGACTGACCAACTGCCTTCGAGTATTAGCCTGTTCCCTGCTTTGGATAGCCACAATATGGATAAAGAGGAAAATACAGAATGATAAACAAATCTCCGAAGACATGGAACTGCAAAGGGGTCCATCCGATGATCCAGAACACGAGGACGCTGTTAAATAA

>translation

MEGSTIGMEHSDRNQKTSQTKCCPPRLKIFICALAFCYFSKSMTASYTKSTITQIERRFEIPSSTVGIIDGGFEMGNMLVITLVSYIGAKFHRPKIIGAGVLLMGIGTLLMASPHFIMGRYKYGTAATHTNDAGNFTVISTCSSDSQKTLQQPFSGCRKEEAESSPLWVIVVLGNTMLGIGEASIIPLGMSFVDDYARPENSAFYIGCLNTIKGIGPIFGLVLGSLCANLYVDIGLVNKESVTITPQDSRWVGAWWLGYVVSGLLTVLAAFPFWFLPKALPENSQISLLDNTPQQHKTTPSLTEIVKDFAPTFKRLLTNKIYILCLAYSIVAFNDFAILVTYTPKYLEQQFGQSASKANFLIGVTCVPAVALGVFLSGLMMKRFKWGLLASARVNLLTIVATVFLTVPFFALSCENLDIAGVTVPYQGSTEVQGVTSDVLPSCNADCGCPDLHWDPVCGENGVTYISPCHAGCSSTQGAGRNKTFHDCGCIQSWGLSVGNSSAVLGQCSQNPNCNRMMYLYLGLQSLALFVYSLGAVPLFTMSLRIVDPELKSLSVGVLLLSIRVLGGIPAPIYFGALVDSSCLKWGQRKSCGRGACRIYDIETFRFLFQGLTNCLRVLACSLLWIATIWIKRKIQNDKQISEDMELQRGPSDDPEHEDAVK*

Page 125: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 128 -

Zebrafish Oatp1f2-2 (1725 bp)

>coding sequence ATGGAAGGCAGCACAATTGGCATGGAACATTCAGACAGAAACCAGAAAACATCTCAAACCAAATGCTGCCCTCCAAGGCTGAAGATCTTTATTGCCGCACTTGCCTTCTGTTACTTCTCTAAATCCATGACGGCGTCTTACACCAAAAGCACCATCACTCAGATCGAGAGGCGGTTTGAGATCCCGAGCTCCACCGTTGGCATCATTGATGGGGGTTTCGAAATGGGTAATATGCTGGTCATTACCCTGGTAAGCTATGTTGGCGCTAAGTTTCACAGACCCAAGATCATTGGAGCAGGGGTTCTGCTGATGGGTATAGGAACACTGCTAATGGCTTCACCTCATTTCATTATGGGCCGGTACAAGTATGGTACAGCTGCCACCCACACTAATGATGCTGGTAATTTCACTGTGATCTCTACATGCTCATCCGACTCTCAGAAAACTCTTCAACAGCCTTTTTCTGGATGCCGAAAAGAGGAAGCTGAAAGCTCACCCTTGTGGGTTATAGTGGTTTTGGGAAACACTATGCTTGGGATTGGGGAAGCCAGTATTATCCCTCTGGGAATGTCATTCGTAGATGATTATGCACGGCCAGAAAACTCAGCTTTTTACATTGGTTGTCTGAACACAATTAAAGGGATTGGGCCAATTTTTGGTTTCGTGCTTGGATCTCTTTGTGCTAATCTTTATGTGGATATTGGCCTAGTGAATAAAGAAAGCGTGACCATCACACCACAGGACTCTCGCTGGGTTGGGGCTTGGTGGTTGGGTTATGTGGTGTCTGGCTTGTTGACTGTCCTCGCTGCTTTTCCTTTCTGGTTCTTGCCAAAAGCTCTGCCTGAAAACTCCCAAATCTCACTGCTAGACAACACCCCACAACAGCACAAAACCACACCCAGCCTTACAGAGATAGTCAAAGATTTTGCGCCAACTTTTAAGCGTCTGCTGACCAATAAGATCTACATTCTGTACCTGGCGTACAGTATAGTGGCATTCAACGACTTCGCCATACTTGTAACGTACACGCCAAAGTATCTGGAACAGCAGTTTGGGCAAAGCGCATCCAAAGCCAACTTTCTGATAGACCTCCATTGGGATCCAGTGTGTGGAGAGAATGGAGTGACCTACATCTCCCCTTGCCATGCGGGCTGCAGTTCCACCCAGGGTGCAGGACGGAATAAGACATTCCATGACTGCGGGTGTATTCAGAGCTGGGGACTGAGCGTTGGCAACTCCTCTGCGGTTCTTGGACAGTGTTCACAAAATCCCAACTGCAACAGAATGATGTACCTTTATCTGGGATTGCAGTCTTTGGCCCTCTTTGTGTACAGTCTTGGTGCTGTTCCCCTTTTCACCATGTCTCTGAGGATTGTGGATCCCGAGCTGAAGTCTCTCTCTGTGGGAGTGTTACTGTTATCTATAAGAGTTTTGGGTGGTATTCCTGCTCCCATTTACTTTGGTGCTCTTGTTGACTCCAGCTGTCTGAAATGGGGCCAGAGGAAATCTTGTGGAAGGGGTGCCTGCAGAATTTATGACATTGAGACGTTCAGGTTCCTCTTTCAAGGACTGACCAACTGCCTTCGAGTATTAGCCTGTTCCCTGCTTTGGATAGCCACAATATGGATAAAGAGGAAAATACAGAATGATAAACAAATCTCCGAAGACATGGAACTGCAAAGGGGTCCATCCGATGATCCAGAACACGAGGACGCTGTTAAATAA

>translation

MEGSTIGMEHSDRNQKTSQTKCCPPRLKIFIAALAFCYFSKSMTASYTKSTITQIERRFEIPSSTVGIIDGGFEMGNMLVITLVSYVGAKFHRPKIIGAGVLLMGIGTLLMASPHFIMGRYKYGTAATHTNDAGNFTVISTCSSDSQKTLQQPFSGCRKEEAESSPLWVIVVLGNTMLGIGEASIIPLGMSFVDDYARPENSAFYIGCLNTIKGIGPIFGFVLGSLCANLYVDIGLVNKESVTITPQDSRWVGAWWLGYVVSGLLTVLAAFPFWFLPKALPENSQISLLDNTPQQHKTTPSLTEIVKDFAPTFKRLLTNKIYILYLAYSIVAFNDFAILVTYTPKYLEQQFGQSASKANFLIDLHWDPVCGENGVTYISPCHAGCSSTQGAGRNKTFHDCGCIQSWGLSVGNSSAVLGQCSQNPNCNRMMYLYLGLQSLALFVYSLGAVPLFTMSLRIVDPELKSLSVGVLLLSIRVLGGIPAPIYFGALVDSSCLKWGQRKSCGRGACRIYDIETFRFLFQGLTNCLRVLACSLLWIATIWIKRKIQNDKQISEDMELQRGPSDDPEHEDAVK*

Page 126: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 129 -

Zebrafish Oatp1f3 (1767 bp)

>coding sequence ATGGGTAATATACTGGTCATTACCCTGGTAAGCTATGTTGGCGCTAAGTTTCACAGGCCCAAGATCATTGGAGCAGGGGTTCTGCTGATGGGTATAGGAACACTGCTAATGGCTTCACCTCATTTCATTATGGGCCGGTACAGGTATGGTACAGCTGCCACCCACACTAATGATGCTGGTAATTTCACTGTGATTTCTACATGCTCATCCGACTCTCAGGAAACTCTTCAAAAGCCTTACTCTGGATGCCAACAAGAGGAATCTGGAAGCTCACCTTTGTGGGTGATGGTGCTTTTGGGAAACACTATGCGTGGGATCGGGGAAGCCAGCATTGTCCCTCTGGGAATGTCATTCATAGATGATTATGCACGGCCAGAAAACTCAGCTTTTTACATTGGCTGTCTGAACACACTTGGAGTGATTGGGCCAATTTTCGGTTTCATGCTTGGATCTCTATGTGCTAATCTTTATGTGGATATTGACTCAGTGTATCAAGCAAGCGTGACCATCACACCGCAGGACTATCGCTGGGTTGGGGCTTGGTGGTTGGGTTATGTGGTGTCTGGCCTGTTCACTCTACTTGCTGCTTTTCCTTTGTGGTTCTTTCCAAGAGCTCTACCTGAAAATTCCCAAATTTCTGAGCTAAACCACCCCCGGCAACAGCACAAAACCACCCCCAGCCTTACAGAGATAGTCAAAGATTTTCCACCATCCTTTAAGCATCTACTTACCAACAAGATTTACATTCTGTACCTAGCTCACAGTATTGTGGCATTTAACGACTTTGCCATAGGGGCGTCATACACACCGAAGTATCTGGAACAGCAGTTTGGACAGAGCGCATCCAAAACCAACTTTCTCATAGGAGCGACATCTGTTCCAGGAGTGGCTCTGGGTATTTTTCTGAGTGGGTTGATGATGAAGAGGTTTAAATGGGATCTGCTGACATCTGCAAGAGTGAACTTGTACGCTGGTGTTGCCATGTTGTTTTTGGCCGTACCTTTCTTCGCTCTCAGCTGCGGAAATCTAGAAGTTGCAGGGGTTACAGTGCCCTATCAAGGATCTACTGAAGTTCAAGGTGTAATAGGTGATGTACTCCCTTCTTGCAATGCTGACTGTGGGTGTCCAGACATCCATTGGGATCCAGTGTGTGGAGAGAATGGAGTGACCTACATCTCCCCTTGCCATGCGGGCTGCAATTCCACTCTGGGTGCAGGACGGAATATGACATTCCATGACTGCGGCTGTATTCAGAGCTGGGGACTGAGCGTTGGCAACTCCTCTGCAGTTCTTGGCCAGTGTTCAAGAGATCCCAAATGCAACAGAATGTTCTACATTTTTCTGGGATTGCAGTCTTTGGCCTTCTTTGTGTACAGTCTTGGCACTGTTCCCTATTTCACCTTGTCTTTGAGGATTGTGGACCCTGAGCTGAAGTCAATCTCTGTGGGAATGTTACTGTTCACTGTAAGAGTTTTGGGTGGTATTCCTGCTCCCATTTACTTTGGTGCTCTTATTGACTCTACCTGTCTGAAATGGGGCAGGAATATATCTTGTGGAAGAGGTGCCTGCAGAATTTATGACATCGAGACGTTCAGGTTTCTCTTTCAGGGACTGACCATCGGTCTTCGAGTATTAGCCTATTCCCTCCTCTGGATAGCCACTATAGAGATACAGAGAAAAATGAAGAGTGATAAACAATTGTCCAGAGACACAGAACTCCAAAAGTGTCCATCCAACGATCCAGAACACGAGGAGACTATAAAATAA

>translation

MGNILVITLVSYVGAKFHRPKIIGAGVLLMGIGTLLMASPHFIMGRYRYGTAATHTNDAGNFTVISTCSSDSQETLQKPYSGCQQEESGSSPLWVMVLLGNTMRGIGEASIVPLGMSFIDDYARPENSAFYIGCLNTLGVIGPIFGFMLGSLCANLYVDIDSVYQASVTITPQDYRWVGAWWLGYVVSGLFTLLAAFPLWFFPRALPENSQISELNHPRQQHKTTPSLTEIVKDFPPSFKHLLTNKIYILYLAHSIVAFNDFAIGASYTPKYLEQQFGQSASKTNFLIGATSVPGVALGIFLSGLMMKRFKWDLLTSARVNLYAGVAMLFLAVPFFALSCGNLEVAGVTVPYQGSTEVQGVIGDVLPSCNADCGCPDIHWDPVCGENGVTYISPCHAGCNSTLGAGRNMTFHDCGCIQSWGLSVGNSSAVLGQCSRDPKCNRMFYIFLGLQSLAFFVYSLGTVPYFTLSLRIVDPELKSISVGMLLFTVRVLGGIPAPIYFGALIDSTCLKWGRNISCGRGACRIYDIETFRFLFQGLTIGLRVLAYSLLWIATIEIQRKMKSDKQLSRDTELQKCPSNDPEHEETIK*

Page 127: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 130 -

Zebrafish Oatp1f4-1 (1935 bp)

>coding sequence ATGGAAGGCAGCGCAAACAGGATGGAATATTCAGGCGGAAACCAGAAAACCTCTCAAACCAAATGCTGCTCAAGTCTGAAGATCTTTATCTGCGCACTCGCCTTCAGTTACTTCTCTAAATCCATGACGGGGTCTTACACCAAAAGCACCATCACTCAGATTGAGAAGCGGTTTGAGATCCCGAGCTCCACCGTTGGTTTCATTGATGGCAGCTTCGAAATGGGTAATATGCTGGTCATTACCCTGGTGAGTTACATTGGTGCTAAGTTTCACAGACCCAAGATCATTGGAGCAGGGGTTCTGCTGATGGGTATGGGAACACTGCTAATGGCTTCACCTCATTTCCTTATGGGCCGGTACAAGTATGGTACAGCTGCCACCCACACTAATGATGCTGGTAATATTACTGTGATCTCTACATGCTCATCCGACTCTCAGGAAACTCTTCAACAGCCTTTCTCTGGATGCCAAAAAGAGGAAGCTGAAAGCTCACCCTTCTGGGTGATAGTGGTTTTGGGAAACACTATGCGTGGGATCGGGGAAGCCAGCATCGTCCCTCTGGGAATGTCATTCATAGATGATTATGCACGGCCAGAAAACTCTGCTTTTTACATTGGTTGTCTGAACACTCTCAAAGGGATTGGGCCGATTTTTGGTTACATGCTTGGATCTCTATGTGCTAATCTTTATGTGGATTTTGGCTTAGTGGATCAAGAAAGCGTGACCATCACACCGCAGGACTCTCGCTGGGTTGGGGCTTGGTGGTTGGGTTATGTGGTGTCTGGCTTGTTGACTCTCCTCGCTGCTTTTCCTTTCTGGTTTTTGCCAAAAGTTCTACCTGAAAACTCCCAAATCTCACTGCTAGACAACACCCCACAACAGCACAAAACCACCCCCAGCCTTACAGAGATAGTCAAAGATTTTGCACCAACTTTTAAGCGTCTGCTCACCAATAAGGTCTACATTCTGTACCTGGCGTACAGTATAGTGGCATTCAACGACTTCGCCATAGTTGCAACATACACGCCAAAGTATCTGGAACAGCAGTTTGGGCAAAGTGCATCCAAAGCCAACTTTCTGATAGGAGTGACATCTATTCCGGCAGTAGCGCTGGGTGTTTTCCTGAGTGGGTTGATGATGAAGAGGTTTAAATTGGGTCTGCTGGCATCCGCAAGAGTGAATTTGTTGACCATAGTCGCCACAGTGTTTTTGGCCGTACCTTTCTTCGCTCTCAGCTGTGAAAATCTAGATGTCGCGGGAGTTACAGTGCCCTATCAAAGATCTACTGAAGTTCAAGGTGTAATAAGTGATGTACTCCCTTCCTGCAATGCTGACTGTGGGTGTCCTGACCTCCAATGGGATCCAGTGTGTGGAGAGAATGGAGTGACCTACATCTCCCCTTGCCATGCGGGCTGCAATTCCACCCAGGGTGCAGGACGGAATATGACATTCCATGACTGCAGGTGTATTCAGAGCTGGGGACTGAGCGTTGGCAACTCCTCTGCAGTTCTTGGCCAGTGTTCAAGAGATCCCAACTGCAACAGAATGATCTACATTTATCTGGGATTGCAGTCTTTGGCCTTCTTTTTGTGCGGTCTTGGCATTATTCCCTTTTTCACCATGTCTCTGAGGATTGTGGATCCTGAGCTGAAGTCTCTCTCTGTGGGAATGTTACTGTTAGCTGTAAGAGTTTTGGGTGGTATTCCTGCTCCCATTTACTTTGGTGCTCTTATTGACTCGACCTGTCTGAAATGGGGCCAGAATAAATCGTGTGGAAGAGGTGCCTGCAGAATTTATGACATCGAGACGTTCAGGTTCCCCTTTCAGGGGCTGACCATGTGTCTTCGAGTATTAGCCTGTTCCCTGCTTTGGATAGCCACCATAGAGATAAAGAGAAAAAATACTGAACAACAAACAAATGTCTCAAGACACTGA

>translation

MEGSANRMEYSGGNQKTSQTKCCSSLKIFICALAFSYFSKSMTGSYTKSTITQIEKRFEIPSSTVGFIDGSFEMGNMLVITLVSYIGAKFHRPKIIGAGVLLMGMGTLLMASPHFLMGRYKYGTAATHTNDAGNITVISTCSSDSQETLQQPFSGCQKEEAESSPFWVIVVLGNTMRGIGEASIVPLGMSFIDDYARPENSAFYIGCLNTLKGIGPIFGYMLGSLCANLYVDFGLVDQESVTITPQDSRWVGAWWLGYVVSGLLTLLAAFPFWFLPKVLPENSQISLLDNTPQQHKTTPSLTEIVKDFAPTFKRLLTNKVYILYLAYSIVAFNDFAIVATYTPKYLEQQFGQSASKANFLIGVTSIPAVALGVFLSGLMMKRFKLGLLASARVNLLTIVATVFLAVPFFALSCENLDVAGVTVPYQRSTEVQGVISDVLPSCNADCGCPDLQWDPVCGENGVTYISPCHAGCNSTQGAGRNMTFHDCRCIQSWGLSVGNSSAVLGQCSRDPNCNRMIYIYLGLQSLAFFLCGLGIIPFFTMSLRIVDPELKSLSVGMLLLAVRVLGGIPAPIYFGALIDSTCLKWGQNKSCGRGACRIYDIETFRFPFQGLTMCLRVLACSLLWIATIEIKRKNTEQQTNVSRH*

Page 128: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 131 -

Zebrafish Oatp1f4-2 (1935 bp)

>coding sequence ATGGAAGGCAGCGCAAACAGGATGGAATATTCAGGCGGAAATCAGAAAACCTCTCAAACCAAATGCTGCTCAAGTCTGAAGATCTTTATCTGCGCACTCGCCTTCAGTTACTTCTCTAAATCCATGACGGGGTCTTACACCAAAAGCACCATCACTCAGATTGAGAAGCGGTTTGAGATCCCGAGCTCCACCGTTGGTTTCATTGATGGCAGCTTCGAAATGGGTAATATGCTGGTCATTACCCCGGTGAGTTACATTGGTGCTAAGTTTCACAGACCCAAGATCATTGGAGCAGGGGTTCTGCTGATGGGTATGGGAACACTGCTAATGGCTTCACCTCATTTCCTTATGGGCCGGTACAAGTATGGTACAGCTGCCACCCACACTAATGATGCTGGTAATATTATTGTGATCTCTACATGCTCATCCGACTCTCAGGAAACTCTTCAACAGCCTTTCTCTGGATGCCAAAAAGAGGAAGCTGAAAGCTCACCCTTCTGGGTGATAGTGGTTTTGGGAAACACTATGCGTGGGATCGGGGAAGCCAGCATCGTCCCTCTGGGAATGTCATTCATAGATGATTATGCACGGCCAGAAAACTCTGCTTTTTACATTGGTTGTCTGAACACTCTCAAAGGGATTGGGCCGATTTTTGGTTACATGCTTGGATCTCTATGTGCTAATCTTTATGTGGATTTTGGCTTAGTGGATCAAGAAAGCGTGACCATCACACCGCAGGACTCTCGCTGGGTTGGGGCTTGGTGGTTGGGTTATGTGGTGTCTGGCTTGTTGACTCTCCTCGCTGCTTTTCCTTTCTGGTTTTTGCCAAAAGTTCTACCTGAAAACTCCCAAATCTCATTGCTAGACAACACCCCACAACAGCACAAAACCACCCCCAGCCTTACAGAGATAGTCAAAGATTTTGCACCAGCTTTTAAGCGTCTGCTCACCAATAAGGTCTACATTCTGTACCTGGCGTACAGTATAGTGGCATTCAACGACTTCGCCATAGTTGCAACATACACGCCAAAGTATCTGGAACAGCAGTTTGGGCAAAGTGCATCCAAAGCCAACTTTCTGATAGGAGTGACATCTATTCCGGCAGTAGCGCTGGGTGTTTTCCTGAGTGGGTTGATGATGAAGAGGTTTAAATTGGGTCCGCTGGCATCCGCAAGAGTGAATTTGTTGACCATAGTCGCCACAGTGTTTTTGGCCGTACCTTTCTTCGCTCTCAGCTGTGAAAATCCAGATGTCGCGGGAGTTACAGTGCCCTATCAAAGATCTACTGAAGTTCAAGGTGTAATAAGTGATGTACTCCCTTCCTGCAATGCTGACTGTGGGTGTCCTGACCTCCAATGGGATCCAGTGTGTGGAGAGAATGGAGTGACCTACATCTCCCCTTGCCATGCGGGCTGCAATTCCACCCAGGGTGCAGGACGGAATATGACATTCCATGACTGCAGGTGTATTCAGAGCTGGGGACTGAGCGTTGGCAACTCCTCTGCAGTTCTTGGCCAGTGTTCAAGAGATCCCAACTGCAACAGAATGATCTACATTTATCTGGGATTGCAGTCTTTGGCCTTCTTTTTGTGCGGTCTTGGCATTATTCCCTTTTTCACCATGTCTCTGAGGATTGTGGATCCTGAGCTGAAGTCTCTCTCTGTGGGAATGTTACTGTTAGCTGTAAGAGTTTTGGGTGGTATTCCTGCTCCCATTTACTTTGGTGCTCTTATTGACTCGACCTGTCTGAAATGGGGCCAGAATAAATCGTGTGGAAGAGGTGCCTGCAGAATTTATGACATCGAGACGTTCAGGTTCCTCTTTCAGGGGCTGACCATGTGTCTTCGAGTATTAGCCTGTTCCCTGCTTTGGATAGCCACCATAGAGATAAAGAGAAAAAATACTGAACAACAAACAAATGTCTCAAGACACTGA

>translation

MEGSANRMEYSGGNQKTSQTKCCSSLKIFICALAFSYFSKSMTGSYTKSTITQIEKRFEIPSSTVGFIDGSFEMGNMLVITPVSYIGAKFHRPKIIGAGVLLMGMGTLLMASPHFLMGRYKYGTAATHTNDAGNIIVISTCSSDSQETLQQPFSGCQKEEAESSPFWVIVVLGNTMRGIGEASIVPLGMSFIDDYARPENSAFYIGCLNTLKGIGPIFGYMLGSLCANLYVDFGLVDQESVTITPQDSRWVGAWWLGYVVSGLLTLLAAFPFWFLPKVLPENSQISLLDNTPQQHKTTPSLTEIVKDFAPAFKRLLTNKVYILYLAYSIVAFNDFAIVATYTPKYLEQQFGQSASKANFLIGVTSIPAVALGVFLSGLMMKRFKLGPLASARVNLLTIVATVFLAVPFFALSCENPDVAGVTVPYQRSTEVQGVISDVLPSCNADCGCPDLQWDPVCGENGVTYISPCHAGCNSTQGAGRNMTFHDCRCIQSWGLSVGNSSAVLGQCSRDPNCNRMIYIYLGLQSLAFFLCGLGIIPFFTMSLRIVDPELKSLSVGMLLLAVRVLGGIPAPIYFGALIDSTCLKWGQNKSCGRGACRIYDIETFRFLFQGLTMCLRVLACSLLWIATIEIKRKNTEQQTNVSRH*

Page 129: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 132 -

Zebrafish Oatp2b1 (2034 bp)

>coding sequence ATGACAGCATTCGAACCCCAAGGAAACCCACAGACAGCGCGACCTCGTAGATCAATGGGTCTCTTCAACAATATCAAGTTTTTTGTACTGTGTCATGGGATGCTCCAGCTTTCTCAGCTCCTGGTGTCGGGCTATCTGAAGGCTTCCATCACCACCATCGAGAAAAGATACGGATTTTCCAGTTCAAAGTCAGGACTTTTGGCAGCCTTCAATGAGGTGGGGAACACAGTTCTCATTGTGTTTGTGAGTTTCTTTGGGAGTCGCGTGCATCGACCAAGATGTATTGGAATCGGAGGAATGATTGCAAGCATGGGAGTGTTTCTCATTGCCCTGCCTCACTTCATTAGCGAAAAGTATGAATATACAGAATTCATTAGCAATACCAGACTCAACTCCAGCCTTTGCCAGACACAGGACATAGAATCCAGTCCGACATGCAGTCAAAATGATACAGACCCTCAATTGGGAGTGTATCCAATACTTCTGCTGGGCCAGTTACTGCTGGGCATTGGTGGTGTTCCAATACAACCTTTTGGCATTTCCTACATAGACGATTATGCAGAAAGGAGGAATTCACCCTTCTACCTAGGCATACTTTTTGCAGTTACCGTAATTGGACCAGCTTTTGGCTACATCATGTCATCTGCTGTGCTTCGCTTGTATGTCGACATAGACAAGATGTCTTTAAATGAAAACAAATTAGAACATGGTGACCCCAGATGGATTGGTGCTTGGTGGTTGGGATTCCTGATAGCTGCCACCCTTCTTTCCTTGACATCTTTGCCATACCTGTTTTTCCCTAGAAAGATGTCCAAAGAGGAGATCGAAGAAGCCCCTGTGGAGCCATCAACTGAGAAAATGATACAAAACAAGAAGCCCGAAACTGACCAGATTGAAGAGGTGTCTCTTACACAGTTCTTAAAAAGCTTTCCCACAATTGTTCTAAGGACTCTGCGAAACCCAATTTATCTGCTGGTAGTGTTCGCTCAGGTGAATCTTGCTGCCATGGTTGCGGGACTGGCAACATTTATGGGCAAATTCATAGAGAAGCAGTTTGCACGGACTGCCTCCTTTTCCAATATCATGATGGGTGGTATCAATATTCCTTCAGCCATGTTTGGCATAATGGCTGGAGGGGTTATTCTACGCAGACTGGGTTTGACTGTCAGGTCCTCTGCGGCGATGTGTACGATAGCAGTCTTCATAAGCATTATGTTTGCGATTCCACTTCTTTTTATTGGCTGCCCCACACAAAAGATTTCAGGACTCAACTACCGCGATTCTCAGCAGTGCAGTAATTCGTGTTACTGTTCTGATGAAGCCTTCAACCCCGTGTGTGGCTCAGATGGTGTGGAGTTTCGCTCTCCCTGTCATGCGGGATGTAAAATAGTCAACATAAAAACAATAGGCAGGATACAGAACGTGACCTACACAGACTGTGGATGTGTGAAATCCAGTGGTTCTCCAGGCTTTGCTTTTGCTGGACCATGTGCAAGTGGCTGTAGTAATCTCCTCATTCCCTTTATGATTCTGTCCTCCCTCACTTGCTTTGTGGCTTCACTTTCCCACACTACATCCTTCATGATGATCCTCAGGACCGTGACTCCGGAAGATAAGTCTTTTGCCGTGGGAATTCAGTTCATGCTGTTCAGAGTTCTTGCTTTCCTCCCATCCCCAGTGCTGTATGGTACTGCGATCGACACTGCATGCCTTGTGTGGGGAAAGAAATGCAGTAAATCAACATCATGTCTGTACTATAATTTAGACCTTTTCAGACACAGATTCCTGGGACTCCAGATATTTTTTATATGTGGTGGCTTCGTCTGTTTCCTTCTGTCCTTTTTGGTGCTTCGGAGGGCAGATTCATCTCAACAACAACAACATCAACAACAACAACAACAACAACAACAACAACAACAACAACAACATCATGAAATGAACATTCAAACCAAAAGTGAAAGTTCCCAGGAAACGAAAGATGATAAACAGACTGTGAGCAATTCAAATGGACACAGCAATACAACGTGTCCTTAG

>translation

MTAFEPQGNPQTARPRRSMGLFNNIKFFVLCHGMLQLSQLLVSGYLKASITTIEKRYGFSSSKSGLLAAFNEVGNTVLIVFVSFFGSRVHRPRCIGIGGMIASMGVFLIALPHFISEKYEYTEFISNTRLNSSLCQTQDIESSPTCSQNDTDPQLGVYPILLLGQLLLGIGGVPIQPFGISYIDDYAERRNSPFYLGILFAVTVIGPAFGYIMSSAVLRLYVDIDKMSLNENKLEHGDPRWIGAWWLGFLIAATLLSLTSLPYLFFPRKMSKEEIEEAPVEPSTEKMIQNKKPETDQIEEVSLTQFLKSFPTIVLRTLRNPIYLLVVFAQVNLAAMVAGLATFMGKFIEKQFARTASFSNIMMGGINIPSAMFGIMAGGVILRRLGLTVRSSAAMCTIAVFISIMFAIPLLFIGCPTQKISGLNYRDSQQCSNSCYCSDEAFNPVCGSDGVEFRSPCHAGCKIVNIKTIGRIQNVTYTDCGCVKSSGSPGFAFAGPCASGCSNLLIPFMILSSLTCFVASLSHTTSFMMILRTVTPEDKSFAVGIQFMLFRVLAFLPSPVLYGTAIDTACLVWGKKCSKSTSCLYYNLDLFRHRFLGLQIFFICGGFVCFLLSFLVLRRADSSQQQQHQQQQQQQQQQQQQQHHEMNIQTKSESSQETKDDKQTVSNSNGHSNTTCP*

Page 130: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

- 133 -

Zebrafish Oatp4a1-1 (2229 bp)

>coding sequence ATGCCCCACCTGTTGAACAATGACACTTCCTTCAGCTCCAAGCAAGAGCTTCTGGTCCTACATAACAGCCCATTGGGACCCAGTCCAGACACTCCCAGTCCTGTGGAGTCCCAGGGCTCTGGTTCAGGCAGCCCGTCTGGAGATGCCCACACTCCACTCCACAATACAGGCAAATTAGGATTGGAGCTCACAGATCCCCCAAATGGGTTGAAGTTTGTCCCAACAGACTCTGAGAATCTGTGTGGATGGGGAGCACTGACCCCGCGGTGTATCCAAACATTCAACACCCCACGATGGGTTCTGTTCTTCCTCTGTGTGGCCTCCTTCCTCCAGGGCATGATCATCAACGGCTTCATCAACACCGTCATCACCTCAATCGAAAGACGTTTTGACCTGCGCAGCTACCAGGCAGGACTGATTGCAAGCTCCTATGACATTGCTGCTTGTTTGTGCCTGACATTTGTGAGCTACTTTGGTGGTACTGGTCATAAACCTCGATGGCTGGGTTGGGGTGTGCTCGTCGTGGCTCTGGGCTCTCTGGTCTTTGCCCTGCCCCATTTCACCACACCAGCCTATCAGGTGAGAGTTCCAGAGCGCATGGGACTGTGCTCAGAAAATCACACAGAAACGTGCGTTGGCGAGGAAGGCGGGCTCTCGGCATACCGCTATGTGTTTATGCTGGGACAGTTTCTTCATGGTGTTGGGGCAACCCCACTATACACCCTTGGGGTCACATACCTTGATGAGAACGTCAAGTCAAACTACGCACCAGTATATATAGGGATCTTCTACACAGCCGCCATCGTAGGACCTGCGGCCGGATACCTGCTGGGTGGATTCTTTCTCAACATTTACACTGAGATTGGCCAAACGACAGAACTGACACCTGAAAACCCGTTGTGGGTTGGAGCTTGGTGGATTGGTTTCCTGGCAGGAGGAGCGGCAGCTCTGCTTATCGCTCTGCCTATTCTGGGTTACCCACGGCAGCTGCCAGGCTCTCAGCGGTATGTGGCTATGCGCGTCTCAGAGGCTCACCAGCTAAAGGACGGCAGTCAGGTCACAGCATCTGATCCTCAATCTGGAAAAACTGTACAAGACATGCCCAGGTCAATGCTGCTTCTACTAAAGAACCCTACCTTCATCTTTCTGTGTCTTGCTGGTGCTACGGAGGCCACGCTGATAGCTGGGATGTCCACATTTGGCCCCAAATTTCTTGAGTCTCAGTTTAGTCTGAGTGCTTCAGAGGCAGCAACCTGGTTTGGGTATATGGTAGTGCCAGCTGGTGGAGGTGGCACCTTTCTGGGTGGCTTCATTATAAAGAAGCTGAATCTTCGCTGTCGTGGGATCATCCGTTTCTGCATGCTGTGTGCACTGGTCAGTCTGATGGCTATCTTTATCTTTTTGGTGCACTGTCCTAATGTGCCTATGGCTGGAGTAACCATCTCATACTACAGCAACGTAACACACAACCACAACAGCGCCCTCTACCTGCAGAAAGACAAACACAACAGCAGCTTCTCAGATCTGAGCAACCTGACCGTAGCCTGTAACATTGGATGTCACTGTCTAGAGGAGTCATTTAACCCTGTGCGTGGTGCAGATGGAGTGATGTATTTCTCTCCCTGTCATGCCGGCTGCAGTTCCCTCAACCACACATTCGGTCCGAGAGGCAGACAGGTGTTCTCTGGCTGCAGCTGTGTGGCTGGGAATATATCATGGGGAGAGCAGGGTTTTGCTGAAGAAGGGAGATGTGTGTCGTCCTGCAATCACATGCCGGCCTTCCTCACCTTCCTCTTCATCCTCATCCTTTTCACTTTCCTCTGCAGCATCCCTGCACTTACTGCCACACTGAGGTGTGTTCCTGACAGTCAGAGATCATTTGGACTGGGAATCCAGTGGATCGTGGTCAGGACTCTGGGTGGGATCCCAGGCCCCATAGCGTTCGGCTCTGTGATCGACATCTCCTGCCTGCTGTGGGAGGAGCAGTGTGGGGAATACGGCTCCTGTTACCTGTACCACAACTCAGCCATGAGTCAGTACTCGCTCGTCGCCGGCATCATCTACAAGGTTTTAGGCACTTTTTTCTTCTTGCTGGCCACCCTGCTTTACAAGCCGCCCCCTGAGTCTCCTCAGAGCAGCTGTGAGAGCACAGATGTAGGAGAGAGCAAAGACCTCCCCATCAAAGACCACCCTGCAGAAATCATCTCCAACCCACACGCTAAATTATGA

>translation

MPHLLNNDTSFSSKQELLVLHNSPLGPSPDTPSPVESQGSGSGSPSGDAHTPLHNTGKLGLELTDPPNGLKFVPTDSENLCGWGALTPRCIQTFNTPRWVLFFLCVASFLQGMIINGFINTVITSIERRFDLRSYQAGLIASSYDIAACLCLTFVSYFGGTGHKPRWLGWGVLVVALGSLVFALPHFTTPAYQVRVPERMGLCSENHTETCVGEEGGLSAYRYVFMLGQFLHGVGATPLYTLGVTYLDENVKSNYAPVYIGIFYTAAIVGPAAGYLLGGFFLNIYTEIGQTTELTPENPLWVGAWWIGFLAGGAAALLIALPILGYPRQLPGSQRYVAMRVSEAHQLKDGSQVTASDPQSGKTVQDMPRSMLLLLKNPTFIFLCLAGATEATLIAGMSTFGPKFLESQFSLSASEAATWFGYMVVPAGGGGTFLGGFIIKKLNLRCRGIIRFCMLCALVSLMAIFIFLVHCPNVPMAGVTISYYSNVTHNHNSALYLQKDKHNSSFSDLSNLTVACNIGCHCLEESFNPVRGADGVMYFSPCHAGCSSLNHTFGPRGRQVFSGCSCVAGNISWGEQGFAEEGRCVSSCNHMPAFLTFLFILILFTFLCSIPALTATLRCVPDSQRSFGLGIQWIVVRTLGGIPGPIAFGSVIDISCLLWEEQCGEYGSCYLYHNSAMSQYSLVAGIIYKVLGTFFFLLATLLYKPPPESPQSSCESTDVGESKDLPIKDHPAEIISNPHAKL*

Zebrafish Oatp4a1-2 (2229 bp)

>coding sequence ATGCCCCACCTGTTGAACAATGACACTTCCTTCAGCTCCAAGCAAGAGCTTCTGGTCCTACATAACAGCCCATTGGGACCCAGTCCAGACACTCCCAGTCCTGTGGAGTCCCAGGGCTCTGGTTCAGGCAGCCCGTCTGGAGATGCCCACACTCCACTCCACAATACAGGCAAATTAGGATTGGAGCTCACAGATCCCCCAAATGGGTTGAAGTTTGTC

Page 131: Molecular cloning, characterization and relevance for

SUPPLEMENTAL MATERIAL Novel Sequences identified in the present study

134

CCAACAGACTCTGAGAATCTGTGTGGATGGGGAGCACTGACCCCGCGGTGTATCCAAACATTCAACACCCCACGATGGGTTCTGTTCTTCCTCTGTGTGGCCTCCTTCCTCCAGGGCATGATCATCAACGGCTTCATCAACACCGTCATCACCTCAATCGAAAGACGTCTTGACCTGCGCAGCTACCAGGCAGGACTGATTGCAAGCTCCTATGACATTGCTGCTTGTTTGTGCCTGACATTTGTGAGCTACTTTGGTGGTACTGGTCATAAACCTCGATGGCTGGGTTGGGGTGTGCTCGTCATGGCTCTGGGCTCTCTGGTCTTTGCCCTGCCCCATTTCACCACACCAGCCTATCAGGTGAGAGTTCCAGAGCGCATGGGACTGTGCTCAGAAAATCACACAGAGACGTGCGTTGGCGAGGAAGGCGGGCTCTCGGCATACCGCTATGTGTTTATGCTGGGACAGTCTCTTCATGGTGTTGGGGCAACCCCACTATACACCCTTGGGGTCACATACCTTGATGAGAACGTCAAGTCAAACTACGCACCAGTATATATAGGGATCTTCTACACAGCCGCCATCGTAGGACCTGCGGCCGGATACCTGCTGGGTGGATTCTTTCTCAACATTTACACTGAGATTGGCCAAACGACAGAACTGACACCTGAAAACCCGTTGTGGGTTGGAGCTTGGTGGATTGGTTTCCTGGCAGGAGGAGCGGCAGCTCTGCTTATCGCTCTGCCTATTCTGGGTTACCCACGGCAGCTGCCAGGCTCTCAGCGGTATGTGGCTATGCGCGTCTCAGAGGCTCACCAGCTAAAGGACGGCAGTCAGGTCACAGCATCTGATCCTCAATTTGGAAAAACTGTAAAAGACATGCCCAGGTCAATGCTGCTTCTACTAAAGAACCCTACCTTCATCTTTCTGTGTCTTGCTGGTGCTACGGAGGCCACGCTGATAGCTGGGATGTCCACATTTGGCCCCAAATTTCTTGAATCTCAGTTTAGTCTGAGTGCTTCAGAGGCAGCAACCTGGTTTGGGTATATGGTAGTGCCAGCTGGTGGAGGTGGCACCTTTCTGGGTGGCTTCATTATAAAGAAGCTGAATCTTCGCTGTCGTGGGATCATCCGTTTCTGCATGCTGTGTGCACTGGTCAGTCTGATGGCTATCTTTATCTTTTTGGTGCACTGTCCTAATGTGCCTATGGCTGGAGTAACCATCTCATACTACAGCAACGTAACACACAACCACAACAGCGCCCTCTACCTGCAGAAAGACAAACACAACAGCAGCTTCTCAGATCTGAGCAACCCGACCGTAGCCTGTAACATTGGATGTCACTGTCTAGAGGAGTCATTTAACCCTGTGTGTGGTGCAGATGGAGTGATGTATTTCTCTCCCTGTCATGCCGGCTGCAGTTCCCTCAACCACACATTCGGTCCGAGAGGCAGACAGGTGTTCTCTGGCTGCAGCTGTGTGGCTGGGAATATATCATGGGGAGAGCAGGGTTTTGCTGAAGAAGGGAGATGTGTGTCGTCCTGCAATCACATGCCAGCCTTCCTCACCTTCCTCTTCATCCTCATCCTTTTCACTTTCCTCTGCAGCATCCCTGCACTTACTGCCACACTGAGGTGTGTTCCTGACAGTCAGAGATCATTTGGACTGGGAATCCAGTGGATCGTGGTCAGGACTCTGGGTGGGATCCCAGGCCCCATAGCGTTCGGCTCTGTGATCGACATCTCCTGCCTGCTGTGGGAGGAGCAGTGTGGGGAATACGGCTCCTGTTACCTGTACCACAACTCAGCCATGAGTCAGTACTCGCTCGTCGCCGGCATCATCTACAAGGTTTTAGGCACTTTTTTCTTCTTGCTGGCCACCCTGCTTTACAAGCCGCCCCCTGAGTCTCCTCAGAGCAGCTGTGAGAGCACAGATGTAGGAGAGAGCAAAGACCTCCCCATCAAAGACCACCCTGCAGAAATCATCTCCAACCCACACGCTAAATTATGA

>translation

MPHLLNNDTSFSSKQELLVLHNSPLGPSPDTPSPVESQGSGSGSPSGDAHTPLHNTGKLGLELTDPPNGLKFVPTDSENLCGWGALTPRCIQTFNTPRWVLFFLCVASFLQGMIINGFINTVITSIERRLDLRSYQAGLIASSYDIAACLCLTFVSYFGGTGHKPRWLGWGVLVMALGSLVFALPHFTTPAYQVRVPERMGLCSENHTETCVGEEGGLSAYRYVFMLGQSLHGVGATPLYTLGVTYLDENVKSNYAPVYIGIFYTAAIVGPAAGYLLGGFFLNIYTEIGQTTELTPENPLWVGAWWIGFLAGGAAALLIALPILGYPRQLPGSQRYVAMRVSEAHQLKDGSQVTASDPQFGKTVKDMPRSMLLLLKNPTFIFLCLAGATEATLIAGMSTFGPKFLESQFSLSASEAATWFGYMVVPAGGGGTFLGGFIIKKLNLRCRGIIRFCMLCALVSLMAIFIFLVHCPNVPMAGVTISYYSNVTHNHNSALYLQKDKHNSSFSDLSNPTVACNIGCHCLEESFNPVCGADGVMYFSPCHAGCSSLNHTFGPRGRQVFSGCSCVAGNISWGEQGFAEEGRCVSSCNHMPAFLTFLFILILFTFLCSIPALTATLRCVPDSQRSFGLGIQWIVVRTLGGIPGPIAFGSVIDISCLLWEEQCGEYGSCYLYHNSAMSQYSLVAGIIYKVLGTFFFLLATLLYKPPPESPQSSCESTDVGESKDLPIKDHPAEIISNPHAKL

Page 132: Molecular cloning, characterization and relevance for

CHAPTER 7

MISCELLANEOUS

Page 133: Molecular cloning, characterization and relevance for

MISCELLANEOUS List of figures

137

LIST OF FIGURES Figure 1: Structure of Microcystin-LR. ............................................................................................................................ - 6 -

Figure 2: Topological model of human OATP1B1. ........................................................................................................ - 16 -

Figure 3: Phylogenetic tree of selected vertebrates in the OATP1 family. ................................................................... - 30 -

Figure 4: Organ distribution of rtOatp1d1. ................................................................................................................... - 31 -

Figure 5: Functional characterization of rtOatp1d1 in HEK293 cells. ........................................................................... - 32 -

Figure 6: Kinetic analyses of rtOatp1d1-mediated substrate uptake. .......................................................................... - 33 -

Figure 7: MC-LR uptake mediated by rtOatp1d1. ......................................................................................................... - 35 -

Figure 8: Competitive inhibition of rtOatp1d1-mediated uptake of MC-LR by TCA and BSP. ...................................... - 35 -

Figure 9: Phylogenetic analysis of representative organic anion transporting polypeptides (OATPs/Oatps) from

mammals and fish. ............................................................................................................................................... - 50 -

Figure 10: Phylogenetic analysis of representative mammalian and fish Oatps belonging to the family 1 ................. - 51 -

Figure 11: Amino acid alignment of full length Oatp1d1 sequences in several fish species ........................................ - 52 -

Figure 12: Expression-levels of Oatp1d1-mRNA in various organs ............................................................................... - 54 -

Figure 13: Global alignments of several zebrafish Oatps cloned in this study. ............................................................ - 69 -

Figure 14: Transport of Oatp substrates. ..................................................................................................................... - 71 -

Figure 15: zfOatp1d1 and zfOatp1f-subtype mediated MC transport. ........................................................................ - 72 -

Figure 16: Quantification of time dependent MC uptake by zfOatps via densitometry analysis. ................................ - 73 -

Figure 17: Concentration dependent uptake of MC-LR and MC-LF (10 nM-1000 nM) mediated by zfOatp1d1 using

immunoblotting with anti-ADDA antibody. ......................................................................................................... - 74 -

Figure 18: RNA expression of (A) zfOatp1d1 and (B) zfOatp1f using semi-quantitative real time PCR. ....................... - 75 -

Page 134: Molecular cloning, characterization and relevance for

MISCELLANEOUS List of Tables

- 138 -

LIST OF TABLES Table 1: Affinities for the tested substrates in this study compared to human and fish transporter of the Oatp1 family. -

37 -

Table 2: Amino acid identities [%] of the Alignment .................................................................................................... - 53 -

Table 3: Variations of cloned amino acid sequences compared to annotated sequences. .......................................... - 68 -

Table 4: Gene lenght and location of human and fish OATPs/Oatps belonging to family 1. ........................................ - 84 -

Supplemental Table 1: Accession numbers for sequences used in phylogenetic tree analysis ................................. - 115 -

Supplemental Table 2: rtOatp1d1 specific primer pairs used for PCR and real time PCR. ......................................... - 116 -

Supplemental Table 3: Accession numbers for sequences used in phylogenetic tree analysis. ................................ - 116 -

Supplemental Table 4: rtOatp1d1 specific primer pairs used for semi quantitative real time PCR ........................... - 118 -

Supplemental Table 5: Annotated Oatp sequence used for primer design ............................................................... - 119 -

Supplemental Table 6: Accession numbers of the zebrafish Oatp subtype sequences ............................................. - 119 -

Supplemental Table 7: Predicted number of transmembrane domains .................................................................... - 120 -

Page 135: Molecular cloning, characterization and relevance for

MISCELLANEOUS Abbreviations

- 139 -

ABBREVIATIONS [3H] Tritium °C Grad Celcius µ micro (10−6) 32PdCTP deoxycytdine triphophate containing the radioactive isotope phosphorus 32 A ampere ABC ATP binding cassette Adda (2S,3S,8S,9S)-3-amino-9-methoxy-2,6,8-trimethyl-10-phenyldeca-4,6-dienoic acid b bases BCA bicinchoninic acid BMAA β-methylamino-L-alanine bp base pairs BSP bromosulfophthalein bw body weight CaCl2 Calciumchloride CAT catalase cDNA compementary Deoxyribonucleic acid Ci Curie CO2 carbon dioxide ct threshold cycle CYS cysteine Da Dalton DHEAS dehydroepiandrosterone sulfate sodium salt DIG Digoxigenin DMEM Dulbecco's Modified Eagle Medium dNTP deoxyribonucleotide DPDPE D-penicillamine 2,5]encephalin DTT Dithiothreitol dw dry weight e.g. for example E17βG estradiol 17-β-glucuronide E3S estrone-3-sulfate EC50 median effective concentration ECL enhanced chemiluminescence EDTA Ethylenediaminetetraacetic acid ef1α elongation factor alpha equ. Equivalent EST elongated sequence tag F Phenylalanine FBS fetalbovine serum g gramm GIT gastrointestinal tract GPx glutathione-peroxidase GR glutathione-reductase GSH glutathione GSSG Glutathion-Disulfid h hours HCl hydrogen chloride HEK293 human embryonic kidney cells Hepes 4-(2-Hydroxyethyl)piperazine-1-ethanesulfonic acid HIV human immunodeficiency virus HPLC high perfomance liquid chromatography HRP horseradish peroxidase IC50 median inhibitory concentration IMAC interim maximum acceptable concentration ip intraperitoneal k kilo KCl Kaliumchloride Km Michaeliskonstante L Leucine l liter LC50 median lethal concentration LCMS liquid chromatography-mass spectrometry LD50 median lethal dose MC Microcystin m milli M Molar Mdha N-methyldehydroalanine Mdr multidrug resistant proteins (P-glycoproteins) MeAsp D-erythro-β-methylaspartic acid MgCl2 magnesiumchloride min minute mRNA messenger ribonucleic acid Mrp multidrug resistance associated proteins MTX methotrexate n nano (10−9) NaCl sodiumchloride NCBI National celter for Biotechnology Information NTCP Na+/taurocholate cotransporting polypeptide Oatp organic anion transporting polypeptides ORF open reading frame p pico (10−12) PCR polymerase chain reaction PKA/PKC protein kinase A/C PP protein phosphatase qRT quantitative real time RACE rapid amplification of cDNA ends RFU relative fluorescence units RNA ribonucleic acid ROS reactive oxygen species rRNA ribosomal ribonucleic acid rt rainbow trout SDS sodium dodecyl sulfate s second SEM standard error of the mean Ser serine SLC solute carrier SLCO solute carrier organic anion transporting family SNP single nucleotide polymorphism SOD superoxide dismutase SSC saline-sodium-citrate T3 triiodthronin T4 thyroxin TBq Tera Becquerel TCA taurocholic acid TDI tolerable daily intake Thr threonine TMD transmembrane domains U unit UTR untranslated region W Tryptophan WHO World Health Organization ww wet weight Δ delta

Page 136: Molecular cloning, characterization and relevance for

MISCELLANEOUS Declaration of self-contribution

- 141 -

DECLARATION OF SELF-CONTRIBUTION This work was conducted without the help of unauthorized third parties, Data and concepts from

other sources discussed in this work are referenced. This work, or a similar was not yet presented to

another examination authority in any country. This thesis was written solely by K.S. The rough concept

of this thesis was part of DFG grant written by the project coordinator D.R.D. and formed into definite

projects by K.S. under the supervision of D.R.D and B.H.. All experiments were designed in

consultation with D.R.D. at the University of Konstanz, unless stated otherwise. The self-contribution

to the manuscripts are as stated below.

Chapter 2 – Molecular cloning and functional characterization of a rainbow trout liver Oatp

K.S. designed and performed all experiments and analyzed all data. Constructing the cDNA library and

experimental design of first characterization experiments (the concept of cell transfection and uptake

experiments) was done in consultation with B.H. in Kansas-City, KS, USA. K.S. wrote the manuscript, all

authors revised the final version of the manuscript.

Chapter 3 – Molecular cloning and organ distribution of Oatp1d1 in whitefish and common carp

K.S. designed all experiments and analyzed all data. Constructing of a cDNA library of common carp

liver and organ distribution of Oatp1d1 in carp organs was performed by H.B. during her master thesis

under the supervision of K.S.. All other experiments were performed by K.S. K.S. wrote the manuscript,

B.H. revised the version of the manuscript presented in this work.

Chapter 4 – Oatp mediated transport of Microcystin-congener in Danio rerio

K.S. designed all experiments. Performing of experiments and analysis of data was done by K.S. and L.Z.

during her master thesis under the supervision of K.S. Figures for the manuscript were partly

compiled by L.Z. and reworked by K.S. K.S. wrote the manuscript, all authors revised the version of the

manuscript presented in this work.

Page 137: Molecular cloning, characterization and relevance for

MISCELLANEOUS Acknowledgements

- 143 -

ACKNOWLEDGEMENTS This work would not have been possible without the support of many people, to those I want to

express my deep gratitude:

Prof. Daniel Dietrich for welcoming me into his group, for providing ideas, for believing in me and

supporting me, for making me feel everything will work out fine, but also for challenging me in ways

that made me outgrow myself into thinking like a scientist, and of course for the wonderful

opportunities to work abroad.

Prof. Bruno Hagenbuch for co-supervising my project without hesitation, allowing me the use of his

laboratory and thereby giving me an additional inspiring work environment, for training me in

relevant techniques and for the numerous skype conversations which made so many things clearer.

The whole group Human- and Environmental toxicity for creating a great atmosphere, for being

colleagues beyond work and for giving me this funny new nickname. Special mention to Julia

Kleinteich and Philipp Secker who always had time to discuss experimental problems and who

became real friends.

The IMPRS for allowing me the opportunity to constantly learn new things in many workshops and

become part of this great, young scientific surroundings. Special mention to Daniel Piechowski and

Mäggi Hieber Ruiz, the coordinators and heart of the school, who always had time when they were

needed.

My thesis committee for discussing my project regularly, for providing new ideas and getting me

back on track more than once.

My students, Steffen Hörer, Fatima Baker, Tim Karner, Julia Kemmler, Bettina Hodapp and

Susanne Ermert who taught me, with their one million questions, to be a more patient person and for

contributing to establishment of methods and first experiments. Especially my master students Lisa

Zimmermann and Heinke Bastek for making my life so much easier with their ambition and

kindness and for making me feel like a proud mom.

Jonathan Puddick and Susie Wood for providing a wonderful work environment and time in Nelson,

New Zealand, for showing me that constructive work practices need not to be based on stress, for

refining my field work practices and for proofreading chapters of my thesis.

David Gustav for helping me clarify the fog of thoughts in this thesis by chain sawing and ping-pong

and tetris playing, without any hope of gratitude.

Page 138: Molecular cloning, characterization and relevance for

MISCELLANEOUS Acknowledgements

- 144 -

My friends Paul Szyszka, Sarah Koch, Henne Kusche, Claudi Fitz and Jacob Stierle for making

Konstanz an unforgettable time in my life, for providing me with delicious food so many times, for

being my chill out zone and for always listening to my problems or just helping me to forget them for a

moment. Mauricio Vargas Uribe, Mary Ashley Rimmer and Linda Mufich for doing the same in the

US.

Robbie Siataga for always believing in me during the writing of this thesis, for providing a shoulder to

moan on and for not being scared of my demons.

The climbing crew in Konstanz, the tango community in Nelson and the hang-gliding folks in

France for igniting passions beyond science.

Meiner ganzen Familie für ihre uneingeschränkte Liebe, besonders Gabriele Steiner die in all den

Jahren sicherlich mein größter Motivator gewesen ist und mich auf allen erdenklichen Ebenen

unterstützt und immer an mich geglaubt hat und Hanne Steiner, die stets bemüht war, Gott für mich

ins Boot zu holen.

Page 139: Molecular cloning, characterization and relevance for

MISCELLANEOUS Summary

- 145 -

SUMMARY In the present dissertation project, transport proteins belonging to the solute carrier family were

identified in several fish species and characterized regarding their organ specific expression and

substrate specificity including the uptake of the cyanotoxin Microcystin (MC).

MC is one of many toxins which are built by Cyanobacteria (blue-green algae). These prokaryotes can

mainly be found in fresh water, but they also inhabit marine waters or even several extreme habitats.

Under favorable conditions, rapid growth of cyanobacteria leads to a mass occurrence (algae blooms),

which is accompanied by high concentrations of toxins within the cyanobacterial cells and also in the

water after lysis of the cells. As a result, Cyanobacteria pose a threat to organisms of several

development stages, trophic levels and habitats. Especially fish are exposed to cyanobacterial toxins,

as they share the same habitat and thus have limited possibilities of avoidance. Indeed, MC

accumulations could be demonstrated in numerous field- and laboratory experiments. As a

consequence of the MC accumulation, pathological alterations of organs and physiology or even death

of fish were observed. Interestingly, the sensitivity towards MC seems to depend on the fish species,

the reasons for that however remain unclear.

MC is the most current cyanotoxins worldwide. More than 100 congeners have been described so far

of the 1000 Da heptapeptide. These mainly exhibit differences in their amino acid composition, but

also in their toxicity. Because of the first-pass-effect, MC uptake mainly results in cytotoxicity of the

liver, but also kidney- and neurotoxicity was found. The cytotoxicity is caused by a functional

inhibition of Serine- and Threonine Protein Phosphatases, which is mediated by covalent binding of

MC to these enzymes. This leads to a misbalance of the phosphorylation state in the cell, which results

in cytoskeletal disintegration, altered signal transduction pathways and finally to apoptosis and cell

necrosis. These toxicodynamics require the uptake of the toxin into the target cell. Because of the

structure and chemical properties of MC, a carrier mediated transport is necessary. In humans and

rodents it could be shown that this transport is mediated by the cell membrane protein Organic anion

transporting polypeptide (Oatp). Also in the little skate (Rajas erinacea), a cartilaginous fish, Oatp

mediated uptake of the MC congener MC-LR could be verified.

Several subtypes of Oatps exist in more than 40 species, which can be classified into families and

subfamilies according to their amino acid identity. In general, they are able to transport a wide range

of endo- and xenobiotic substances and they can be expressed in several organs, although subtype

specific differences were found.

The uptake of the toxin into the cell displays a critical crux for the kinetics and consequently the

dynamics and the resulting cytotoxicity. To estimate, which organs are targets for MC intoxication and

Page 140: Molecular cloning, characterization and relevance for

MISCELLANEOUS Summary

- 146 -

which fish species are likely to be more sensitive towards MC, it is crucial to identify possible MC

transporting Oatps and investigate their organ expression profile in different fish species.

Based on that, goal of the study was

• The cloning of novel Oatps from different native fish species, inhabiting the Lake Constance,

the identification of the organ specific expression profile of these Oatps and their functional

characterization.

• The identification of Oatps in native and model fish species, which are able to transport MC.

• The comparison of the uptake capacities of different MC-congeners mediated by several Oatp

subtypes.

We identified and cloned an Oatp sequence using a cDNA library from the rainbow trout

(Oncorhynchus mykiss) liver and classified it as Oatp1d1. This transporter was so far only found in fish.

To functional characterize the novel Oatp1d1 transporter, we transiently transfected it into HEK293-

cells. With that we could investigate the transport of radioactive labelled substances and the kinetics

of chosen substrates. We exposed transfected cells to MC to indirectly demonstrate the uptake of the

toxin by the novel transporter using immunoblotting. Functional characterization showed that

Oatp1d1 is multispecific, similar to the human transporters OATP1B1, OATP1B3 and OATP1A2 and is

expressed predominately in the liver like OATP1B1 and OATP1B3. As opposed to OATP1B1 and

OATP1B3, Oatp1d1 could additionally be found in the brain, what could be the reason for MC mediated

neurotoxicity.

By comparing the organ expression profile of Oatp1d1 in two fish species, found in the Lake Constance

we aimed to get a better understanding about the species specific sensitivity towards MC. Therefore

we identified and cloned this transporter from the liver of the salmonid whitefish (Coregonus

wartmanni) and the cyprinid common carp (Cyprinius carpio) using RACE-PCR or a cDNA library,

respectively. Using semi-quantitative real time PCR we could show that Oatp1d1 in whitefish is mainly

expressed in the liver and to some extend in the brain. A similar Oatp1d1 expression was found in

rainbow trout, a closely related salmonid. In comparison to that, an additional Oatp1d1 expression

was found in the kidney of carp. From these results we draw the conclusion that Oatp1d1 is one

transporter responsible for the MC uptake in the liver and probably also in the brain in several fish

species. Additionally this transporter could participate in transporting MC into kidney cells in carp.

Nevertheless, MC accumulations were also observed in other organs, which cannot be explained by the

expression of Oatp1d1. Therefore the expression of other Oatps capable of MC uptake can be assumed.

As several Oatps are annotated and published in the model organism zebrafish (Danio rerio), their

Page 141: Molecular cloning, characterization and relevance for

MISCELLANEOUS Summary

- 147 -

sequence was used to screen further Oatp transporters regarding their MC uptake. For that we used

several MC-congeners in order to investigate differences in their toxicity. Additionally to Oatp1d1 we

could identify Oatp1f2 and Oatp1f4 as MC transporter. Oatp1d1 transported all tested congeners (MC-

LR, MC-LW, MC-LF and MC-RR), while Oatp1f2 only mediated the transport of MC-LR, MC-LW and MC-

LF. Both transporters however exhibited a higher uptake velocity of MC-LW and MC-LF in comparison

to MC-LR. This leads to the conclusion that lower concentrations of MC-LW und MC-LF are necessary

in comparison to MC-LR to exhibit the same dynamics in fish. Oatp1f4 only transported MC-LF, even

though Oatp1f2 and Oatp1f4 have an amino acid identity of 87 %. This indicates that differences in

single amino acids can contribute to the substrate specificity of the transporter.

Taken together, we could identify novel Oatp sequences in fish, which are contributing to MC transport

in several organs and thereby play a key role in the kinetics of the toxin. We could demonstrate the

transport of several MC-congeners mediated by these transporters. Moreover the expression of

Oatp1d1 depends on the species investigated, what could explain species specific differences in MC

sensitivity. Thus, these results contribute to the clarification of the toxicokinetics of MC in fish, which is

essential for assessing the risk this toxin has on the ecosystem but also human health.

Page 142: Molecular cloning, characterization and relevance for

MISCELLANEOUS Zusammenfassung

- 149 -

ZUSAMMENFASSUNG In der vorliegenden Arbeit wurden Transportproteine der Solute Carrier Family in verschiedenen

Fischspezies identifiziert. Diese Transporter wurden charakterisiert in Bezug auf deren Organ

spezifische Expression und außerdem deren Substratspezifität, wobei der Fokus auf dem Transport

des Cyanotoxins Microcystin (MC) lag.

MC ist eines der vielen Toxine das von Cyanobakterien (auch Blaualgen genannt) gebildet wird. Diese

kommen weltweit hauptsächlich im Süßwasser vor, konnten aber auch in marinen Gewässern

gefunden werden. Unter günstigen Bedingungen findet ein Massenauftreten (sogenannte Algenblüten)

statt und damit einhergehend eine hohe Produktion an Toxinen, die nach Zersetzung der Zellen in das

Wasser abgegeben werden. Aus diesem Grund stellen Cyanobakterien eine Gefahr für Organismen

verschiedener Entwicklungsstufen, trophischer Ebenen und Habitate dar. Da Fische einer toxischen

Cyanobakterienblüte nur bedingt ausweichen können, sind sie besonders exponiert. Tatsächlich

wurde MC in zahlreichen Feld- und Laborexperimenten in verschiedenen Fischspezies nachgewiesen.

Nach akuter MC- Exposition wurden pathologischen Organveränderungen bis hin zu Massensterben

von Fischen beobachtet. Interessanterweise weisen dabei verschiedene Fischarten eine

unterschiedliche Sensitivität gegenüber MC auf, deren Ursache bisher ungeklärt ist. MC ist das

weltweit am häufigsten vorkommende Cyanotoxin. Es handelt sich um ein ca. 1000 Da schweres

Heptapeptid. Bisher sind mehr als 100 Kongenere dieses Toxins bekannt, welche sich hauptsächlich in

einzelnen Aminosäuren voneinander unterscheiden, aber auch in ihrer Toxizität differieren

Microcystine erzeugen aufgrund des first-pass-effects eine hohe Zytotoxizität in der Leber, weisen

jedoch auch eine markante Nieren- und Neurotoxizität auf. Die Zytotoxizität wird durch die kovalente

Bindung von MC an Serin- und Threonin Protein Phosphatasen und somit deren funktionellen

Inhibition vermittelt. Die daraus resultierende Veränderung des Phosphorylierungsgleichgewichts der

Zelle führt zu Zerfall des Zytoskeletts, veränderter Signaltransduktion, oxidativem Stress und

schließlich Zelltod durch Apoptose und Nekrose. Um Phosphatasen inhibieren zu können, muss MC

zunächst allerdings zunächst in die Zelle gelangen. Aufgrund der Struktur, Größe und chemischen

Eigenschaften von MC ist eine Aufnahme mittels Membrantransportern notwendig. Im Menschen und

in Nagern konnte gezeigt werden, dass dieser Transport durch das Zellmembranprotein Organic Anion

transporting polypeptides (Oatp) vermittelt wird. Auch im Rochen (Rajas erinacea), einem

Knorpelfisch, konnte der Transport des MC-Kongeners MC-LR durch Oatp nachgewiesen werden.

Es gibt diverse Subtypen von Oatps in mehr als 40 Spezies, welche auf Grund der Ähnlichkeit ihrer

Aminosäuresequenz in Familien und Subfamilien eingeteilt werden können. Generell transportieren

Oatps ein weites Spektrum an endo- und xenobiotischen Substraten und können in allen Organen

vorkommen, jedoch gibt es subtypenspezifische Unterschiede.

Page 143: Molecular cloning, characterization and relevance for

MISCELLANEOUS Zusammenfassung

- 150 -

Die Aufnahme eines Toxins in die Zelle stellt eine kritische Schlüsselstelle in der Kinetik und die

dadurch vermittelte Dynamik und der finalen Zytotoxizität dar. Durch Identifizierung neuer Oatp-

Subtypen in verschiedenen Fischarten und deren Expressionsmuster in verschiedenen Fischorganen

kann somit abgeschätzt werden, welche Organe von einer MC Intoxikation betroffen sein könnten und

welche Fischarten sensibler gegenüber MC sind.

Ziel dieser Arbeit war daher

• Die Klonierung neuer Oatps aus verschiedenen einheimischen Fischarten, welche im Bodensee

vorkommen sowie die Charakterisierung dieser Oatps bezüglich organspezifischer Expression

und Funktionalität.

• Die Identifizierung von Oatps in einheimischen und Modell-Fischen, welche zu einem MC

Transport beitragen können.

• Die Rolle von verschiedenen Oatps in Zebrafischen bei der Kongener abhängigen

Toxikokinetik.

Dazu wurde zunächst eine Oatp Sequenz mittels einer cDNA-Bibliothek aus der Leber der

Regenbogenforelle (Oncorhynchus mykiss) kloniert, welche als Oatp1d1 klassifiziert wurde. Es handelt

sich hierbei um einen Subtyp, der bisher nur in Fischen gefunden wurde. Der Transport von

bekannten Oatp Substraten wurde daraufhin mit radioaktiv markierten Substanzen untersucht, von

ausgewählten Substanzen wurde zusätzlich eine Kinetik erstellt. Um zu überprüfen ob es sich um

einen MC-Transporter handelt wurden Oatp transfizierte HEK293-Zellen MC ausgesetzt und die

Aufnahme des Toxins indirekt mittels Immunoblot nachgewiesen. Die funktionelle Charakterisierung

zeigte, dass es sich um einen multispezifischen Transporter handelt, der ähnlich wie die im Menschen

vorhandenen Transporter OATP1B3 und OATP1B1 diverse Substrate inklusive MC transportiert und

primär in der Leber vorkommt. Im Gegensatz zu OATP1B3 und OATP1B1 konnte Oatp1d1 jedoch auch

im Gehirn der Fische gefunden werden, was eine Neurotoxizität von MC begründen könnte.

Der Vergleich des Expressionsmuster von Oatps zweier im Bodensee lebenden Fischarten sollte

Aufschluss über die Sensitivität gegenüber MC in verschiedenen Spezies geben. Dafür wurde der

fischspezifische Oatp1d1, welcher sowohl im Rochen als auch in der Regenbogenforelle MC

transportiert, aus der Leber des Bodenseefelchens (Coregonus wartmanni, Salmonidae) und des

Karpfens (Cyprinius carpio, Cyprinidae) kloniert. Hierbei wurde für das Karpfen Oatp eine cDNA-

Bibliothek erstellt, die Oatp1d1 Sequenz aus dem Felchen wurde mittels RACE-PCR isoliert. Mittels

semi-quantitativer real time PCR konnte gezeigt werden, dass Oatp1d1 im Felchen hauptsächlich in

der Leber exprimiert wird, jedoch kann auch im Gehirn Oatp1d1 RNA nachgewiesen werden. Eine fast

identische Organverteilung von Oatp1d1 konnte zuvor in der nah verwandten Regenbogenforelle

Page 144: Molecular cloning, characterization and relevance for

MISCELLANEOUS Zusammenfassung

- 151 -

gefunden werden. Im Vergleich dazu zeigt sich im Karpfen eine zusätzliche Expression in der Niere,

welche quantitativ zwischen Gehirn und Leber liegt. Aus diesen Ergebnissen geht hervor, dass

Oatp1d1 ein MC Transporter in verschiedenen Fischspezies in der Leber und höchstwahrscheinlich

auch im Gehirn ist. Zusätzlich könnte er zu einem Transport in die Niere im Karpfen beitragen.

Dennoch kann durch die Expression von Oatp1d1 noch nicht die hauptsächlich in Karpfen und

Regenbogenforelle beobachtete Aufnahme von MC in andere Organe erklärt werden. Es liegt nahe, das

weitere Transporter vorhanden sein müssen, welche für die Aufnahme von MC verantwortlich sind. Da

für den Modellorganismus Zebrafisch (Danio rerio) alle Oatp Subtypen publiziert und annotiert sind,

wurde deren Sequenz genutzt um weitere Transporter bezügliche ihrer MC Aufnahme zu untersuchen.

Dabei wurden verschiedene Kongenere angeschaut um die beobachteten Unterschiede in der

vermittelten Toxizität erklären zu können. Es konnte gezeigt werden, dass neben Oatp1d1 auch die

nierenspezifischen Transporter Oatp1f2 und Oatp1f4 MC transportieren. Während Oatp1d1 alle

getesteten MC-Kongenere (MC-LR, MC-LW, MC-LF und MC-RR) transportiert, transportiert Oatp1f2

nur MC-LR, MC-LW und MC-LF. Beide Transporter zeigten jedoch eine höhere

Aufnahmegeschwindigkeit für MC-LW und MC-LF im Vergleich zu MC-LR, was deren höhere bzw.

schneller vermittelte Toxizität begründet.

Obwohl Oatp1f2 und Oatp1f4 in ihrer Aminosäuresequenz zu ca. 87 % identisch sind, transportiert

Oatp1f4 nur MC-LF. Dies deutet darauf hin, dass Unterschiede in einzelnen Aminosäuren zur

Substratspezifität beitragen können.

Zusammenfassend konnten durch diese Dissertation neue Oatp-Sequenzen in Fischen gefunden

werden, die für den MC-Transport in verschiedene Organe verantwortlich sind und daher eine

Schlüsselrolle in der Kinetik von MC darstellen. Es konnte gezeigt werden, dass diverse MC-Kongenere

transportiert werden und die Expression der Oatps von der jeweiligen Fischspezies abhängt, was eine

individuelle Speziessensitivität erklären könnte. Diese Ergebnisse unterstützen daher die Aufklärung

der Toxikokinetik von MC in Fischen, die zur Risikoabschätzung sowohl für die Auswirkungen auf das

Ökosystem als auch auf die menschliche Gesundheit essentiell sind.

Page 145: Molecular cloning, characterization and relevance for

CHAPTER 8

REFERENCES

Page 146: Molecular cloning, characterization and relevance for

Abe, T., Kakyo, M., Tokui, T., Nakagomi, R., Nishio, T., Nakai, D., Nomura, H., Unno, M., Suzuki, M., and Naitoh, T. (1999). Identification of a novel gene family encoding human liver-specific organic anion transporter LST-1. Journal of Biological Chemistry 274(24), 17159-17163.

Abe, T., Unno, M., Onogawa, T., Tokui, T., Kondo, T. N., Nakagomi, R., Adachi, H., Fujiwara, K., Okabe, M., and Suzuki, T. (2001). LST-2, a human liver-specific organic anion transporter, determines methotrexate sensitivity in gastrointestinal cancers. Gastroenterology 120(7), 1689-1699.

Adamovský, O., Kopp, R., Hilscherová, K., Babica, P., Palíková, M., Pašková, V., Navrátil, S., Maršálek, B., and Bláha, L. (2007). Microcystin kinetics (bioaccumulation and elimination) and biochemical responses in common carp (Cyprinus carpio) and silver carp (Hypophthalmichthys molitrix) exposed to toxic cyanobacterial blooms. Environmental Toxicology and Chemistry 26(12), 2687-2693.

Albay, M., Akcaalan, R., Tufekci, H., Metcalf, J. S., Beattie, K. A., and Codd, G. A. (2003). Depth profiles of cyanobacterial hepatotoxins (microcystins) in three Turkish freshwater lakes. Hydrobiologia 505(1-3), 89-95.

Amado, L., Garcia, M., Pereira, T., Yunes, J., Bogo, M., and Monserrat, J. (2011). Chemoprotection of lipoic acid against microcystin-induced toxicosis in common carp (Cyprinus carpio, Cyprinidae). Comparative biochemistry and physiology. Toxicology & pharmacology: CBP 154(3), 146-153.

Amrani, A., Nasri, H., Azzouz, A., Kadi, Y., and Bouaïcha, N. (2014). Variation in Cyanobacterial Hepatotoxin (Microcystin) Content of Water Samples and Two Species of Fishes Collected from a Shallow Lake in Algeria. Archives of environmental contamination and toxicology 66(3), 379-389.

An, J., and Carmichael, W. W. (1994). Use of a colorimetric protein phosphatase inhibition assay and enzyme linked immunosorbent assay for the study of microcystins and nodularins. Toxicon 32(12), 1495-1507.

Andersen, R. J., Luu, H. A., Chen, D. Z., Holmes, C. F., Kent, M. L., Le Blanc, M., Taylor, F. M., and Williams, D. E. (1993). Chemical and biological evidence links microcystins to salmon ‘netpen liver disease’. Toxicon 31(10), 1315-1323.

Badagnani, I., Castro, R. A., Taylor, T. R., Brett, C. M., Huang, C. C., Stryke, D., Kawamoto, M., Johns, S. J., Ferrin, T. E., and Carlson, E. J. (2006). Interaction of methotrexate with organic-anion transporting polypeptide 1A2 and its genetic variants. Journal of Pharmacology and Experimental Therapeutics 318(2), 521-529.

Baganz, D., Staaks, G., Pflugmacher, S., and Steinberg, C. E. (2004). Comparative study of microcystin-LR-induced behavioral changes of two fish species, Danio rerio and Leucaspius delineatus. Environmental toxicology 19(6), 564-570.

Baganz, D., Staaks, G., and Steinberg, C. (1998). Impact of the cyanobacteria toxin, microcystin-LR on behaviour of zebrafish, danio rerio. Water research 32(3), 948-952.

Bartram, J., Carmichael, W. W., Chorus, I., Jones, G., and Skulberg, O. M. (1999). Eutrophication, cyanobacterial blooms and surface scums. In Toxic cyanobacteria in water: A guide to their public health consequences, monitoring and management (I. Chorus, and J. Bartram),chapter 1.2, pp. 3-4. Spon Press.

Bergwerk, A., Shi, X., Ford, A., Kanai, N., Jacquemin, E., Burk, R., Bai, S., Novikoff, P., Stieger, B., and Meier, P. (1996). Immunologic distribution of an organic anion transport protein in rat liver and kidney. American Journal of Physiology-Gastrointestinal and Liver Physiology 271(2), G231-G238.

Page 147: Molecular cloning, characterization and relevance for

REFERENCES

- 156 -

Best, J., Eddy, F., and Codd, G. (2001). Effects of purified microcystin-LR and cell extracts of Microcystis strains PCC 7813 and CYA 43 on cardiac function in brown trout (Salmo trutta) alevins. Fish Physiology and Biochemistry 24(3), 171-178.

Boaru, D. A., Dragoş, N., and Schirmer, K. (2006). Microcystin-LR induced cellular effects in mammalian and fish primary hepatocyte cultures and cell lines: a comparative study. Toxicology 218(2), 134-148.

Bossuyt, X., Müller, M., and Meier, P. J. (1996). Multispecific amphipathic substrate transport by an organic anion transporter of human liver. Journal of hepatology 25(5), 733-738.

Botes, D., Kruger, H., and Viljoen, C. (1982). Isolation and characterization of four toxins from the blue-green alga, Microcystis aeruginosa. Toxicon: official journal of the International Society on Toxinology 20(6), 945.

Botes, D. P., Tuinman, A. A., Wessels, P. L., Viljoen, C. C., Kruger, H., Williams, D. H., Santikarn, S., Smith, R. J., and Hammond, S. J. (1984). The structure of cyanoginosin-LA, a cyclic heptapeptide toxin from the cyanobacterium Microcystis aeruginosa. Journal of the Chemical Society, Perkin Transactions 1, 2311-2318.

Briz, O., Romero, M. R., Martinez-Becerra, P., Macias, R. I., Perez, M. J., Jimenez, F., San Martin, F. G., and Marin, J. J. (2006). Oatp8/1B3-mediated cotransport of bile acids and glutathione. An export pathway for organic anions from hepatocytes? Journal of Biological Chemistry.

Brooks, W., and Codd, G. (1987). Distribution of Microcystis aeruginosa peptide toxin and interactions with hepatic microsomes in mice. Pharmacology & toxicology 60(3), 187-191.

Bürgi, H., and Stadelmann, P. (2002). Change of phytoplankton composition and biodiversity in Lake Sempach before and during restoration. Hydrobiologia 469(1-3), 33-48.

Bury, N., Eddy, F., and Codd, G. (1995). The effects of the cyanobacterium Microcystis aeruginosa, the cyanobacterial hepatotoxin microcystin–LR, and ammonia on growth rate and ionic regulation of brown trout. Journal of Fish Biology 46(6), 1042-1054.

Bury, N., Flik, G., Eddy, F., and Codd, G. (1996). The effects of cyanobacteria and the cyanobacterial toxin microcystin-LR on Ca2+ transport and Na+/K+-ATPase in tilapia gills. The Journal of experimental biology 199(6), 1319-1326.

Cai, S.-Y., Wang, W., Soroka, C. J., Ballatori, N., and Boyer, J. L. (2002). An evolutionarily ancient Oatp: insights into conserved functional domains of these proteins. American Journal of Physiology-Gastrointestinal and Liver Physiology 282(4), G702-G710.

Campos, A., and Vasconcelos, V. (2010). Molecular mechanisms of microcystin toxicity in animal cells. International journal of molecular sciences 11(1), 268-287.

Canfield, D. E. (2005). The early history of atmospheric oxygen: homage to Robert M. Garrels. Annual Review of Earth and Planetary Sciences 33, 1-36.

Carbis, C., Rawlin, G., Mitchell, G., Anderson, J., and McCauley, I. (1996). The histopathology of carp, Cyprinus carpio L., exposed to microcystins by gavage, immersion and intraperitoneal administration. Journal of Fish Diseases 19(3), 199-207.

Carmichael, W., He, J.-w., Eschedor, J., He, Z.-r., and Juan, Y.-M. (1988). Partial structural determination of hepatotoxic peptides from Microcystis aeruginosa(cyanobacterium) collected in ponds of central China. Toxicon 26(12), 1213-1217.

Page 148: Molecular cloning, characterization and relevance for

REFERENCES

- 157 -

Carmichael, W. W. (2001). Health effects of toxin-producing cyanobacteria:“The CyanoHABs”. Human and ecological risk assessment: An International Journal 7(5), 1393-1407.

Cattori, V., Hagenbuch, B., Hagenbuch, N., Stieger, B., Ha, R., Winterhalter, K. E., and Meier, P. J. (2000). Identification of organic anion transporting polypeptide 4 (Oatp4) as a major full-length isoform of the liver-specific transporter-1 (rlst-1) in rat liver. FEBS letters 474(2), 242-245.

Cazenave, J., Bistoni, M. d. l. A., Pesce, S. F., and Wunderlin, D. A. (2006). Differential detoxification and antioxidant response in diverse organs of Corydoras paleatus experimentally exposed to microcystin-RR. Aquatic Toxicology 76(1), 1-12.

Cazenave, J., Nores, M., Miceli, M., Díaz, M., Wunderlin, D., and Bistoni, M. (2008). Changes in the swimming activity and the glutathione S-transferase activity of Jenynsia multidentata fed with microcystin-RR. Water research 42(4-5), 1299-1307.

Cazenave, J., Wunderlin, D., de Los Angeles, B. M., Amé, M., Krause, E., Pflugmacher, S., and Wiegand, C. (2005). Uptake, tissue distribution and accumulation of microcystin-RR in Corydoras paleatus, Jenynsia multidentata and Odontesthes bonariensis. A field and laboratory study. Aquatic Toxicology 75(2), 178.

Chen, J., and Xie, P. (2005). Seasonal dynamics of the hepatotoxic microcystins in various organs of four freshwater bivalves from the large eutrophic Lake Taihu of subtropical China and the risk to human consumption. Environmental toxicology 20(6), 572-584.

Chen, J., Zhang, D., Xie, P., Wang, Q., and Ma, Z. (2009). Simultaneous determination of microcystin contaminations in various vertebrates (fish, turtle, duck and water bird) from a large eutrophic Chinese lake, Lake Taihu, with toxic Microcystis blooms. Science of the Total Environment.

Chen, Y., Zeng, S.-F., and Cao, Y.-F. (2012). Oxidative stress response in zebrafish (Danio rerio) gill experimentally exposed to subchronic microcystin-LR. Environmental monitoring and assessment 184(11), 6775-6787.

Choudhuri, S., Ogura, K., and Klaassen, C. D. (2000). Cloning of the full-length coding sequence of rat liver-specific organic anion transporter-1 (rlst-1) and a splice variant and partial characterization of the rat lst-1 gene. Biochemical and biophysical research communications 274(1), 79-86.

Christoffersen, K. (1996). Ecological implications of cyanobacterial toxins in aquatic food webs. Phycologia 35(6S), 42-50.

Cox, P. A., Banack, S. A., and Murch, S. J. (2003). Biomagnification of cyanobacterial neurotoxins and neurodegenerative disease among the Chamorro people of Guam. Proceedings of the National Academy of Sciences 100(23), 13380-13383.

Craig, M., and Holmes, C. F. B. (2000). Freshwater Hepatotoxins: Microcystin and Nodularin, Mechanisms of Toxicity and Effects on Health. In Seafood and freshwater toxins pharmacology, physiology and detection (L. M. Botana),chapter 30, pp. 643-672. CRC Press, USA.

Craig, M., Luu, H. A., McCready, T. L., Holmes, C. F., Williams, D., and Andersen, R. J. (1996). Molecular mechanisms underlying the interaction of motuporin and microcystins with type-1 and type-2A protein phosphatases. Biochemistry and cell biology 74(4), 569-578.

Cui, Y., König, J., Leier, I., Buchholz, U., and Keppler, D. (2001). Hepatic uptake of bilirubin and its conjugates by the human organic anion transporter SLC21A6. Journal of Biological Chemistry 276(13), 9626-9630.

Page 149: Molecular cloning, characterization and relevance for

REFERENCES

- 158 -

Davies, A., Blakeley, A. G. H., and Kidd, C. (2001). Human Physiology. Churchill Livingstone.

Deblois, C. P., Aranda-Rodriguez, R., Giani, A., and Bird, D. F. (2008). Microcystin accumulation in liver and muscle of tilapia in two large Brazilian hydroelectric reservoirs. Toxicon 51(3), 435-448.

Deblois, C. P., Giani, A., and Bird, D. F. (2011). Experimental model of microcystin accumulation in the liver of Oreochromis niloticus exposed subchronically to a toxic bloom of Microcystis sp. Aquatic Toxicology 103(1), 63-70.

Dietrich, D., Fischer, A., Michel, C., and Hoeger, S. (2008). Toxin mixture in cyanobacterial blooms-a critical comparison of reality with current procedures employed in human health risk assessment. Advances in Experimental Medicine and Biology 619, 885.

Dietrich, D., and Hoeger, S. (2005). Guidance values for microcystins in water and cyanobacterial supplement products (blue-green algal supplements): a reasonable or misguided approach? Toxicology and applied pharmacology 203(3), 273-289.

Dong, G., Xie, S., Zhu, X., Han, D., Yang, Y., Song, L., Gan, L., and Chen, W. (2012). Responses of yellow catfish (Pelteobagrus fulvidraco Richardson) exposed to dietary cyanobacteria and subsequent recovery. Toxicon: official journal of the International Society on Toxinology 60(7), 1298.

Engström-Öst, J., Lehtiniemi, M., Green, S., Kozlowsky-Suzuki, B., and Viitasalo, M. (2002). Does cyanobacterial toxin accumulate in mysid shrimps and fish via copepods? Journal of Experimental Marine Biology and Ecology 276(1), 95-107.

Eriksson, J., Toivola, D., Meriluoto, J., Karaki, H., Han, Y., and Hartshorne, D. (1990). Hepatocyte deformation induced by cyanobacterial toxins reflects inhibition of protein phosphatases. Biochemical and biophysical research communications 173(3), 1347-1353.

Ernst, B., Hitzfeld, B., and Dietrich, D. (2001). Presence of Planktothrix sp. and cyanobacterial toxins in Lake Ammersee, Germany and their impact on whitefish (Coregonus lavaretus L.). Environmental toxicology 16(6), 483-488.

Ernst, B., Hoeger, S., O'Brien, E., and Dietrich, D. (2006). Oral toxicity of the microcystin-containing cyanobacterium Planktothrix rubescens in European whitefish (Coregonus lavaretus). Aquatic Toxicology 79(1), 31.

Ernst, B., Hoeger, S. J., O’Brien, E., and Dietrich, D. R. (2007). Physiological stress and pathology in European whitefish (Coregonus lavaretus) induced by subchronic exposure to environmentally relevant densities of Planktothrix rubescens. Aquatic Toxicology 82, 15-26.

Ernst, B., Hoeger, S. J., O’Brien, E., and Dietrich, D. R. (2009). Abundance and toxicity of Planktothrix rubescens in the pre-alpine Lake Ammersee, Germany. Harmful Algae 8(2), 329-342.

Falconer, I., and Yeung, D. (1992). Cytoskeletal changes in hepatocytes induced by Microcystis toxins and their relation to hyperphosphorylation of cell proteins. Chemico-biological interactions 81(1-2), 181.

Falconer, I. R. (2001). Toxic cyanobacterial bloom problems in Australian waters: risks and impacts on human health. Phycologia 40(3), 228-233.

Falconer, I. R., Bartram, J., Chorus, I., Kuiper-Goodman, T., Utkilen, H., Burch, M., and Codd, G. A. (1999). Safe levels and safe practices. In Toxic cyanobacteria in water: A guide to their public health consequences, monitoring and management (I. Chorus, and J. Bartram),chapter 5, pp. 91-104. Spon Press.

Page 150: Molecular cloning, characterization and relevance for

REFERENCES

- 159 -

Feurstein, D., Holst, K., Fischer, A., and Dietrich, D. R. (2009). Oatp-associated uptake and toxicity of microcystins in primary murine whole brain cells. Toxicology and applied pharmacology 234(2), 247-255.

Feurstein, D., Kleinteich, J., Heussner, A. H., Stemmer, K., and Dietrich, D. R. (2010). Investigation of Microcystin Congener-Dependent Uptake into Primary Murine Neurons Environmental Health Perspectives 118(10), 1370-1375.

Fischer, A., Hoeger, S., Stemmer, K., Feurstein, D., Knobeloch, D., Nussler, A., and Dietrich, D. (2010). The role of organic anion transporting polypeptides (OATPs/SLCOs) in the toxicity of different microcystin congeners in vitro: A comparison of primary human hepatocytes and OATP-transfected HEK293 cells. Toxicology and applied pharmacology 245(1), 9-20.

Fischer, W., and Dietrich, D. (2000). Pathological and Biochemical Characterization of Microcystin-Induced Hepatopancreas and Kidney Damage in Carp (Cyprinus carpio). Toxicology and applied pharmacology 164(1), 73-81.

Fischer, W. J., Altheimer, S., Cattori, V., Meier, P. J., Dietrich, D. R., and Hagenbuch, B. (2005). Organic anion transporting polypeptides expressed in liver and brain mediate uptake of microcystin. Toxicology and applied pharmacology 203(3), 257-263.

Fischer, W. J., Hitzfeld, B. C., Tencalla, F., Eriksson, J. E., Mikhailov, A., and Dietrich, D. R. (2000). Microcystin-LR toxicodynamics, induced pathology, and immunohistochemical localization in livers of blue-green algae exposed rainbow trout (Oncorhynchus mykiss). Toxicological Sciences 54(2), 365-373.

Fournie, J. W., and Courtney, L. A. (2002). Histopathological evidence of regeneration following hepatotoxic effects of the cyanotoxin microcystin-LR in the hardhead catfish and gulf killifish. Journal of Aquatic Animal Health 14(4), 273-280.

Freitas de Magalhães, V., Moraes Soares, R., and Azevedo, S. M. (2001). Microcystin contamination in fish from the Jacarepaguá Lagoon (Rio de Janeiro, Brazil): ecological implication and human health risk. Toxicon 39(7), 1077-1085.

Fricker, G., Dubost, V., Finsterwald, K., and Boyer, J. (1994). Characteristics of bile salt uptake into skate hepatocytes. Biochemical Journal 299(Pt 3), 665-670.

Fricker, G., Hugentobler, G., Meier, P. J., Kurz, G., and Boyer, J. L. (1987). Identification of a single sinusoidal bile salt uptake system in skate liver. American Journal of Physiology - Gastrointestinal and Liver Physiology 253(6), 816-822.

Fu, J., and Xie, P. (2006). The acute effects of microcystin LR on the transcription of nine glutathione S-transferase genes in common carp Cyprinus carpio L. Aquatic Toxicology 80(3), 261-266.

Fujiki, H., and Suganuma, M. (2011). Tumor promoters--microcystin-LR, nodularin and TNF-α and human cancer development. Anti-cancer agents in medicinal chemistry 11(1), 4.

Funari, E., and Testai, E. (2008). Human health risk assessment related to cyanotoxins exposure. CRC Critical Reviews in Toxicology 38(2), 97-125.

Gaete, V., Canelo, E., Lagos, N., and Zambrano, F. (1994). Inhibitory effects of Microcystis aeruginosa toxin on ion pumps of the gill of freshwater fish. Toxicon: official journal of the International Society on Toxinology 32(1), 121.

Gallo, A. (2014), Algae's Return Deals Setback to Lake Erie's Revival. The Wall Street Journal (http://online.wsj.com/articles/algaes-return-deals-setback-to-lake-eries-revival-1407183729)

Page 151: Molecular cloning, characterization and relevance for

REFERENCES

- 160 -

Gkelis, S., Lanaras, T., and Sivonen, K. (2006). The presence of microcystins and other cyanobacterial bioactive peptides in aquatic fauna collected from Greek freshwaters. Aquatic Toxicology 78(1), 32-41.

Gui, C., and Hagenbuch, B. (2009). Role of transmembrane domain 10 for the function of organic anion transporting polypeptide 1B1. Protein Science: A Publication of the Protein Society 18(11), 2298.

Gui, C., Miao, Y., Thompson, L., Wahlgren, B., Mock, M., Stieger, B., and Hagenbuch, B. (2008). Effect of pregnane X receptor ligands on transport mediated by human OATP1B1 and OATP1B3. European journal of pharmacology 584(1), 57-65.

Guryev, V., Koudijs, M. J., Berezikov, E., Johnson, S. L., Plasterk, R. H., Van Eeden, F. J., and Cuppen, E. (2006). Genetic variation in the zebrafish. Genome research 16(4), 491-497.

Hagenbuch, B., and Gui, C. (2008). Xenobiotic transporters of the human organic anion transporting polypeptides (OATP) family. Xenobiotica 38(7-8), 778-801.

Hagenbuch, B., and Meier, P. (2003). The superfamily of organic anion transporting polypeptides. Biochimica et biophysica acta 1609(1), 1.

Hagenbuch, B., and Meier, P. J. (2004). Organic anion transporting polypeptides of the OATP/SLC21 family: phylogenetic classification as OATP/SLCO superfamily, new nomenclature and molecular/functional properties. Pflügers Archiv 447(5), 653-665.

Hagenbuch, B., and Stieger, B. (2013). The SLCO (former SLC21) superfamily of transporters. Molecular aspects of medicine 34(2-3), 396-412.

Hairston, N. G., Lampert, W., Cáceres, C. E., Holtmeier, C. L., Weider, L. J., Gaedke, U., Fischer, J. M., Fox, J. A., and Post, D. M. (1999). Lake ecosystems: rapid evolution revealed by dormant eggs. nature 401(6752), 446-446.

Hänggi, E., Grundschober, A. F., Leuthold, S., Meier, P. J., and St-Pierre, M. V. (2006). Functional analysis of the extracellular cysteine residues in the human organic anion transporting polypeptide, OATP2B1. Molecular pharmacology 70(3), 806-817.

Harada, K.-i. (1995). Chemistry and Detection of Microcystins. In Toxic microcystis (M. F. Watanabe, K.-i. Harada, W. W. Carmichael, and H. Fujiki),chapter 6, pp. 103-148. CRC press.

Hastie, C. J., Borthwick, E. B., Morrison, L. F., Codd, G. A., and Cohen, P. T. (2005). Inhibition of several protein phosphatases by a non-covalently interacting microcystin and a novel cyanobacterial peptide, nostocyclin. Biochimica et Biophysica Acta (BBA)-General Subjects 1726(2), 187-193.

He, J., Chen, J., Xie, P., Zhang, D., Li, G., Wu, L., Zhang, W., Guo, X., and Li, S. (2012). Quantitatively evaluating detoxification of the hepatotoxic microcystins through the glutathione and cysteine pathway in the cyanobacteria-eating bighead carp. Aquatic Toxicology 116, 61-68.

Hermansky, S., Stohs, S., Eldeen, Z., Roche, V., and Mereish, K. (1991). Evaluation of potential chemoprotectants against microcystin-LR hepatotoxicity in mice. Journal of Applied Toxicology 11(1), 65-73.

Heussner, A. H., Altaner, S., Kamp, L., Rubio, F., and Dietrich, D. R. (2014). Pitfalls in microcystin extraction and recovery from human blood serum. Chemico-biological interactions 223(87-94).

Page 152: Molecular cloning, characterization and relevance for

REFERENCES

- 161 -

Höger, S. J., Schmid, D., Blom, J. F., Ernst, B., and Dietrich, D. R. (2007). Analytical and functional characterization of microcystins [Asp3] MC-RR and [Asp3, Dhb7] MC-RR: consequences for risk assessment? Environmental science & technology 41(7), 2609-2616.

Honkanen, R. E., Zwiller, J., Moore, R., Daily, S. L., Khatra, B., Dukelow, M., and Boynton, A. (1990). Characterization of microcystin-LR, a potent inhibitor of type 1 and type 2A protein phosphatases. Journal of Biological Chemistry 265(32), 19401-19404.

Hrudey, S., Burch, M., Drikas, M., and Gregory, R. (1999). Remedial measures. In Toxic cyanobacteria in water: A guide to their public health consequences, monitoring and management (I. Chorus, and J. Bartram),chapter 9, pp. 160-181. Spon Press.

Hsiang, B., Zhu, Y., Wang, Z., Wu, Y., Sasseville, V., Yang, W.-P., and Kirchgessner, T. G. (1999). A novel human hepatic organic anion transporting polypeptide (OATP2) Identification of a liver-specific human organic anion transporting polypeptide and identification of rat and human hydroxymethylglutaryl-CoA reductase inhibitor transporters. Journal of Biological Chemistry 274(52), 37161-37168.

Huang, J., Li, N., Hong, W., Zhan, K., Yu, X., Huang, H., and Hong, M. (2013). Conserved tryptophan residues within putative transmembrane domain 6 affect transport function of organic anion transporting polypeptide 1B1. Molecular pharmacology 84(4), 521-527.

Huber, R., Gao, B., Sidler, P. M., Zhang-Fu, W., Leuthold, S., Hagenbuch, B., Folkers, G., Meier, P., and Stieger, B. (2007). Characterization of two splice variants of human organic anion transporting polypeptide 3A1 isolated from human brain. American journal of physiology. Cell physiology 292(2), C795-806.

Ibelings, B. W., Bruning, K., De Jonge, J., Wolfstein, K., Pires, L. D., Postma, J., and Burger, T. (2005). Distribution of microcystins in a lake foodweb: no evidence for biomagnification. Microbial ecology 49(4), 487-500.

Ibelings, B. W., and Havens, K. E. (2008). Cyanobacterial toxins: a qualitative meta–analysis of concentrations, dosage and effects in freshwater, estuarine and marine biota. Advances in Experimental Medicine and Biology 619, 675-732.

Ito, E., Takai, A., Kondo, F., Masui, H., Imanishi, S., and Harada, K.-i. (2002). Comparison of protein phosphatase inhibitory activity and apparent toxicity of microcystins and related compounds. Toxicon 40(7), 1017-1025.

Jewel, M., Affan, M., and Khan, S. (2003). Fish mortality due to cyanobacterial bloom in an aquaculture pond in Bangladesh. Pakistan Journal of Biological Sciences 6(12), 1046-1050.

Jiang, J., Gu, X., Song, R., Zhang, Q., Geng, J., Wang, X., and Yang, L. (2011). Time-dependent oxidative stress and histopathological changes in Cyprinus carpio L. exposed to microcystin-LR. Ecotoxicology 20(5), 1000-1009.

Jiang, J., Shan, Z., Xu, W., Wang, X., Zhou, J., Kong, D., and Xu, J. (2013). Microcystin-LR Induced Reactive Oxygen Species Mediate Cytoskeletal Disruption and Apoptosis of Hepatocytes in Cyprinus carpio L. PLoS ONE 8(12).

Jones, G., and Orr, P. (1994). Release and degradation of microcystin following algicide treatment of a Microcystis aeruginosa bloom in a recreational lake, as determined by HPLC and protein phosphatase inhibition assay. Water research 28(4), 871-876.

Page 153: Molecular cloning, characterization and relevance for

REFERENCES

- 162 -

Jungblut, A. D., Hawes, I., Mountfort, D., Hitzfeld, B., Dietrich, D. R., Burns, B. P., and Neilan, B. A. (2005). Diversity within cyanobacterial mat communities in variable salinity meltwater ponds of McMurdo Ice Shelf, Antarctica. Environmental microbiology 7(4), 519-529.

Kakyo, M., Sakagami, H., Nishio, T., Nakai, D., Nakagomi, R., Tokui, T., Naitoh, T., Matsuno, S., Abe, T., and Yawo, H. (1999). Immunohistochemical distribution and functional characterization of an organic anion transporting polypeptide 2 (oatp2). FEBS letters 445(2), 343-346.

Kalliokoski, A., and Niemi, M. (2009). Impact of OATP transporters on pharmacokinetics. British journal of pharmacology 158(3), 693-705.

Kangur, K., Kangur, A., Kangur, P., and Laugaste, R. (2005). Fish kill in Lake Peipsi in summer 2002 as a synergistic effect of a cyanobacterial bloom, high temperature, and low water level. In Proceedings of the Estonian Academy of Sciences, Biology, Ecology, pp. 67-80.

Kaplan, A., Harel, M., Kaplan-Levy, R., Hadas, O., Sukenik, A., and Dittmann, E. (2011). The languages spoken in the water body (or the biological role of cyanobacterial toxins). Frontiers in microbiology 3, 138-138.

Kapoor, B., Smit, H., and Verighina, I. (1976). The alimentary canal and digestion in teleosts. Advances in marine biology 13, 109-239.

Karjalainen, M., Reinikainen, M., Spoof, L., Meriluoto, J. A., Sivonen, K., and Viitasalo, M. (2005). Trophic transfer of cyanobacterial toxins from zooplankton to planktivores: consequences for pike larvae and mysid shrimps. Environmental toxicology 20(3), 354-362.

Kent, M., Dawe, S., St. Hilaire, S., and Andersen, R. (1996). Effects of feeding rate, seawater entry, and exposure to natural biota on the severity of net-pen liver disease among pen-reared Atlantic salmon. The Progressive fish-culturist 58(1), 43-46.

Keogh, J. (2011). Membrane transporters in drug development. Advances in pharmacology (San Diego, Calif.) 63, 1-42.

Kohoutek, J., Adamovský, O., Oravec, M., Šimek, Z., Palíková, M., Kopp, R., and Bláha, L. (2010). LC-MS analyses of microcystins in fish tissues overestimate toxin levels—critical comparison with LC-MS/MS. Analytical and bioanalytical chemistry 398(3), 1231-1237.

Komatsu, M., Furukawa, T., Ikeda, R., Takumi, S., Nong, Q., Aoyama, K., Akiyama, S.-i., Keppler, D., and Takeuchi, T. (2007). Involvement of mitogen-activated protein kinase signaling pathways in microcystin-LR–induced apoptosis after its selective uptake mediated by OATP1B1 and OATP1B3. Toxicological Sciences 97(2), 407-416.

Kondo, F., Ikai, Y., Oka, H., Okumura, M., Ishikawa, N., Harada, K., Matsuura, K., Murata, H., and Suzuki, M. (1992). Formation, characterization, and toxicity of the glutathione and cysteine conjugates of toxic heptapeptide microcystins. Chemical research in toxicology 5(5), 591-596.

Kondo, F., Matsumoto, H., Yamada, S., Ishikawa, N., Ito, E., Nagata, S., Ueno, Y., Suzuki, M., and Harada, K.-i. (1996). Detection and identification of metabolites of microcystins formed in vivo in mouse and rat livers. Chemical research in toxicology 9(8), 1355-1359.

König, J. (2011). Uptake transporters of the human OATP family: molecular characteristics, substrates, their role in drug-drug interactions, and functional consequences of polymorphisms. Handbook of experimental pharmacology(201), 1-28.

Page 154: Molecular cloning, characterization and relevance for

REFERENCES

- 163 -

Kotak, B., Semalulu, S., Fritz, D., Prepas, E., Hrudey, S., and Coppock, R. (1996). Hepatic and renal pathology of intraperitoneally administered microcystin-LR in rainbow trout (Oncorhynchus mykiss). Toxicon: official journal of the International Society on Toxinology 34(5), 517-525.

Kozlowsky-Suzuki, B., Wilson, A. E., and Ferrão-Filho, A. d. S. (2012). Biomagnification or biodilution of microcystins in aquatic foodwebs? Meta-analyses of laboratory and field studies. Harmful Algae 18, 47-55.

Krienitz, L., Ballot, A., Kotut, K., Wiegand, C., Pütz, S., Metcalf, J., Codd, G., and Pflugmacher, S. (2003). Contribution of hot spring cyanobacteria to the mysterious deaths of Lesser Flamingos at Lake Bogoria, Kenya. FEMS microbiology ecology 43(2), 141.

Krishnamurthy, T., Carmichael, W., and Sarver, E. (1986). Toxic peptides from freshwater cyanobacteria (blue-green algae). I. Isolation, purification and characterization of peptides from Microcystis aeruginosa and Anabaena flos-aquae. Toxicon 24(9), 865-873.

Kuiper-Goodman, T., Falconer, I. R., and Fitzgerald, J. (1999). Toxicological studies. In Toxic cyanobacteria in water: A guide to their public health consequences, monitoring and management (I. Chorus, and J. Bartram),chapter 4.2, pp. 75-83. Spon Press.

Kullak-Ublick, G. A., Hagenbuch, B., Stieger, B., Schteingart, C. D., Hofmann, A. F., Wolkoff, A. W., and Meier, P. J. (1995). Molecular and functional characterization of an organic anion transporting polypeptide cloned from human liver. Gastroenterology 109(4), 1274-1282.

Kuo, K.-L., Zhu, H., McNamara, P. J., and Leggas, M. (2012). Localization and functional characterization of the rat Oatp4c1 transporter in an in vitro cell system and rat tissues. PLoS ONE 7(6), e39641.

Laemmli, U. K. (1970). Cleavage of structural proteins during the assembly of the head of bacteriophage T4. nature 227(5259), 680-685.

Lawton, L. A., and Codd, G. (1991). Cyanobacterial (blue-green algal) toxins and their significance in UK and European waters. Water and Environment Journal 5(4), 460-465.

Lee, T. K., Koh, A. S., Cui, Z., Pierce, R. H., and Ballatori, N. (2003). N-Glycosylation Controls Functional Activity of Oatp1, an Organic Anion Transporter. American Journal of Physiology - Gastrointestinal and Liver Physiology 285(2), G371-81.

Lee, W., Glaeser, H., Smith, L., Roberts, R., Moeckel, G., Gervasini, G., Leake, B., and Kim, R. (2005). Polymorphisms in human organic anion-transporting polypeptide 1A2 (OATP1A2): implications for altered drug disposition and central nervous system drug entry. The Journal of biological chemistry 280(10), 9610-9617.

Lei, H., Xie, P., Chen, J., Liang, G., Dai, M., and Zhang, X. (2008). Distribution of toxins in various tissues of crucian carp intraperitoneally injected with hepatotoxic microcystins. Environmental toxicology and chemistry/SETAC 27(5), 1167.

Letunic, I., and Bork, P. (2007). Interactive Tree Of Life (iTOL): an online tool for phylogenetic tree display and annotation. Bioinformatics 23(1), 127-128.

Leuthold, S., Hagenbuch, B., Mohebbi, N., Wagner, C. A., Meier, P. J., and Stieger, B. (2009). Mechanisms of pH-gradient driven transport mediated by organic anion polypeptide transporters. American Journal of Physiology-Cell Physiology 296(3), C570-C582.

Page 155: Molecular cloning, characterization and relevance for

REFERENCES

- 164 -

Li, G., Xie, P., Fu, J., Hao, L., Xiong, Q., and Li, H. (2008). Microcystin-induced variations in transcription of GSTs in an omnivorous freshwater fish, goldfish. Aquatic Toxicology 88(1), 75-80.

Li, L., Lee, T. K., Meier, P. J., and Ballatori, N. (1998). Identification of glutathione as a driving force and leukotriene C4 as a substrate for oatp1, the hepatic sinusoidal organic solute transporter. Journal of Biological Chemistry 273(26), 16184-16191.

Li, L., Meier, P., and Ballatori, N. (2000). Oatp2 mediates bidirectional organic solute transport: a role for intracellular glutathione. Molecular pharmacology 58(2), 335.

Li, L., Xie, P., and Chen, J. (2007). Biochemical and ultrastructural changes of the liver and kidney of the phytoplanktivorous silver carp feeding naturally on toxic Microcystis blooms in Taihu Lake, China. Toxicon 49(7), 1042-1053.

Li, L., Xie, P., and Guo, L. (2010). Antioxidant response in liver of the phytoplanktivorous bighead carp (Aristichthys nobilis) intraperitoneally-injected with extracted microcystins. Fish Physiology and Biochemistry 36(2), 165-172.

Li, W., Chen, J., Xie, P., He, J., Guo, X., Tuo, X., Zhang, W., and Wu, L. (2014a). Rapid conversion and reversible conjugation of glutathione detoxification of microcystins in bighead carp (Aristichthys nobilis). Aquatic Toxicology 147, 18-25.

Li, X.-B., Zhang, X., Ju, J., Li, Y., Yin, L., and Pu, Y. (2014b). Alterations in neurobehaviors and inflammation in hippocampus of rats induced by oral administration of microcystin-LR. Environmental Science and Pollution Research, 1-7.

Li, X., Chung, I., Kim, J., and Lee, J. (2004). Subchronic oral toxicity of microcystin in common carp (Cyprinus carpio L.) exposed to Microcystis under laboratory conditions. Toxicon: official journal of the International Society on Toxinology 44(8), 821.

Li, X., Chung, I., Kim, J., and Lee, J. (2005). Oral exposure to Microcystis increases activity-augmented antioxidant enzymes in the liver of loach (Misgurnus mizolepis) and has no effect on lipid peroxidation. Comparative biochemistry and physiology. Toxicology & pharmacology: CBP 141(3), 292-296.

Li, X., Liu, Y., Song, L., and Liu, J. (2003). Responses of antioxidant systems in the hepatocytes of common carp (Cyprinus carpio L.) to the toxicity of microcystin-LR. Toxicon 42(1), 85-89.

Liu, X., Huang, J., Sun, Y., Zhan, K., Zhang, Z., and Hong, M. (2013). Identification of Multiple Binding Sites for Substrate Transport in Bovine Organic Anion Transporting Polypeptide 1a2. Drug metabolism and disposition 41(3), 602-607.

Lovell, R., Schaeffer, D., Hooser, S., Haschek, W., Dahlem, A., Carmichael, W., and Beasley, V. (1989). Toxicity of intraperitoneal doses of microcystin-LR in two strains of male mice. Journal of environmental pathology, toxicology and oncology: official organ of the International Society for Environmental Toxicology and Cancer 9(3), 221.

Lu, H., Choudhuri, S., Ogura, K., Csanaky, I. L., Lei, X., Cheng, X., Song, P.-z., and Klaassen, C. D. (2008). Characterization of organic anion transporting polypeptide 1b2-null mice: essential role in hepatic uptake/toxicity of phalloidin and microcystin-LR. Toxicological Sciences 103(1), 35-45.

Malbrouck, C., and Kestemont, P. (2006). Effects of microcystins on fish. Environmental Toxicology and Chemistry 25(1), 72-86.

Page 156: Molecular cloning, characterization and relevance for

REFERENCES

- 165 -

Malbrouck, C., Trausch, G., Devos, P., and Kestemont, P. (2003). Hepatic accumulation and effects of microcystin-LR on juvenile goldfish Carassius auratus L. Comparative Biochemistry and Physiology, Part C 135(1), 39-48.

Manganelli, M., Scardala, S., Stefanelli, M., Palazzo, F., Funari, E., Vichi, S., Buratti, F. M., and Testai, E. (2012). Emerging health issues of cyanobacterial blooms. Annali dell'Istituto superiore di sanità 48(4), 415-428.

Martins, J. C., and Vasconcelos, V. M. (2009). Microcystin dynamics in aquatic organisms. Journal of Toxicology and Environmental Health, Part B 12(1), 65-82.

Maruyama, T., Kato, K., Yokoyama, A., Tanaka, T., Hiraishi, A., and Park, H.-D. (2003). Dynamics of microcystin-degrading bacteria in mucilage of Microcystis. Microbial ecology 46(2), 279-288.

Masuda, S., Saito, H., Nonoguchi, H., Tomita, K., and Inui, K. (1997). mRNA distribution and membrane localization of the OAT-K1 organic anion transporter in rat renal tubules. FEBS letters 407(2), 127-131.

Maynes, J. T., Luu, H. A., Cherney, M. M., Andersen, R. J., Williams, D., Holmes, C. F., and James, M. N. (2006). Crystal structures of protein phosphatase-1 bound to motuporin and dihydromicrocystin-LA: elucidation of the mechanism of enzyme inhibition by cyanobacterial toxins. Journal of molecular biology 356(1), 111-120.

Meier-Abt, F., Hammann-Hänni, A., Stieger, B., Ballatori, N., and Boyer, J. (2007). The organic anion transport polypeptide 1d1 (Oatp1d1) mediates hepatocellular uptake of phalloidin and microcystin into skate liver. Toxicology and applied pharmacology 218(3), 274-279.

Mikkaichi, T., Suzuki, T., Onogawa, T., Tanemoto, M., Mizutamari, H., Okada, M., Chaki, T., Masuda, S., Tokui, T., and Eto, N. (2004). Isolation and characterization of a digoxin transporter and its rat homologue expressed in the kidney. Proceedings of the National Academy of Sciences of the United States of America 101(10), 3569-3574.

Mohamed, Z. A., and Hussein, A. A. (2006). Depuration of microcystins in tilapia fish exposed to natural populations of toxic cyanobacteria: a laboratory study. Ecotoxicology and environmental safety 63(3), 424-429.

Monks, N. R., Liu, S., Xu, Y., Yu, H., Bendelow, A. S., and Moscow, J. A. (2007). Potent cytotoxicity of the phosphatase inhibitor microcystin LR and microcystin analogues in OATP1B1-and OATP1B3-expressing HeLa cells. Molecular cancer therapeutics 6(2), 587-598.

Moutou, K. A., Tsikogias, S., Papadimitriou, T., and Kagalou, I. (2012). Oxidative stress in Cyprinus carpio to analyze microcystin impact in eutrophic shallow lakes: a preliminary study. Journal of Environmental Monitoring 14(8), 2195-2203.

Mur, L. R., Skulberg, O. M., and Utkilen, H. (1999). Cyanobacteria in the environment. In Toxic cyanobacteria in water: A guide to their public health consequences, monitoring and management (I. Chorus, and J. Bartram),chapter 2, pp. 9-25. Spon Press.

Muzzio, A. M., Noyes, P. D., Stapleton, H. M., and Lema, S. C. (2014). Tissue distribution and thyroid hormone effects on mRNA abundance for membrane transporters Mct8, Mct10, and organic anion-transporting polypeptides (Oatps) in a teleost fish. Comparative Biochemistry and Physiology Part A: Molecular & Integrative Physiology 167, 77-89.

Page 157: Molecular cloning, characterization and relevance for

REFERENCES

- 166 -

Nakanishi, T., and Tamai, I. (2012). Genetic polymorphisms of OATP transporters and their impact on intestinal absorption and hepatic disposition of drugs. Drug metabolism and pharmacokinetics 27(1), 106-121.

Neilan, B. A., Pearson, L. A., Muenchhoff, J., Moffitt, M. C., and Dittmann, E. (2013). Environmental conditions that influence toxin biosynthesis in cyanobacteria. Environmental microbiology 15(5), 1239-1253.

Nies, A. T., Niemi, M., Burk, O., Winter, S., Zanger, U. M., Stieger, B., Schwab, M., and Schaeffeler, E. (2013). Genetics is a major determinant of expression of the human hepatic uptake transporter OATP1B1, but not of OATP1B3 and OATP2B1.

Noé, J., Portmann, R., Brun, M.-E., and Funk, C. (2007). Substrate-dependent drug-drug interactions between gemfibrozil, fluvastatin and other organic anion-transporting peptide (OATP) substrates on OATP1B1, OATP2B1, and OATP1B3. Drug metabolism and disposition 35(8), 1308-1314.

Ogura, K., Choudhuri, S., and Klaassen, C. D. (2001). Genomic organization and tissue-specific expression of splice variants of mouse organic anion transporting polypeptide 2. Biochemical and biophysical research communications 281(2), 431-439.

Ohtani, I., Moore, R. E., and Runnegar, M. T. (1992). Cylindrospermopsin: a potent hepatotoxin from the blue-green alga Cylindrospermopsis raciborskii. Journal of the American Chemical Society 114(20), 7941-7942.

Oliver, R. L., and Ganf, G. G. (2000). Freshwater blooms. In The ecology of cyanobacteria: their diversity in time and space (B. A. Whitton, and M. Potts),chapter 6, pp. 149-194. Springer.

Oren, A. (2000). Salts and brines. In The ecology of cyanobacteria: their diversity in time and space (B. A. Whitton, and M. Potts),chapter 10, pp. 281-306. Springer.

Palikova, M., Mares, J., Kopp, R., Hlavkova, J., Navratil, S., Adamovsky, O., Chmelar, L., and Blaha, L. (2011). Accumulation of microcystins in Nile tilapia, Oreochromis niloticus L., and effects of a complex cyanobacterial bloom on the dietetic quality of muscles. Bulletin of environmental contamination and toxicology 87(1), 26-30.

Papadimitriou, T., Kagalou, I., Stalikas, C., Pilidis, G., and Leonardos, I. D. (2012). Assessment of microcystin distribution and biomagnification in tissues of aquatic food web compartments from a shallow lake and evaluation of potential risks to public health. Ecotoxicology 21(4), 1155-1166.

Papadimitriou, T., Katsiapi, M., Kormas, K. A., Moustaka-Gouni, M., and Kagalou, I. (2013). Artificially-born “killer” lake: phytoplankton based water quality and microcystin affected fish in a reconstructed lake. Science of the Total Environment 452, 116-124.

Park, H. D., Sasaki, Y., Maruyama, T., Yanagisawa, E., Hiraishi, A., and Kato, K. (2001). Degradation of the cyanobacterial hepatotoxin microcystin by a new bacterium isolated from a hypertrophic lake. Environmental toxicology 16(4), 337-343.

Pavagadhi, S., Gong, Z., Hande, M. P., Dionysiou, D. D., Armah, A., and Balasubramanian, R. (2012). Biochemical response of diverse organs in adult Danio rerio (zebrafish) exposed to sub-lethal concentrations of microcystin-LR and microcystin-RR: A balneation study. Aquatic Toxicology 109, 1-10.

Page 158: Molecular cloning, characterization and relevance for

REFERENCES

- 167 -

Peng, L., Liu, Y., Chen, W., Liu, L., Kent, M., and Song, L. (2010). Health risks associated with consumption of microcystin-contaminated fish and shellfish in three Chinese lakes: significance for freshwater aquacultures. Ecotoxicology and environmental safety 73(7), 1804-1811.

Pflugmacher, S., Codd, G. A., and Steinberg, C. E. (1999). Effects of the cyanobacterial toxin microcystin-LR on detoxication enzymes in aquatic plants. Environmental toxicology 14(1), 111-115.

Pflugmacher, S., Wiegand, C., Oberemm, A., Beattie, K. A., Krause, E., Codd, G. A., and Steinberg, C. E. (1998). Identification of an enzymatically formed glutathione conjugate of the cyanobacterial hepatotoxin microcystin-LR: the first step of detoxication. Biochimica et Biophysica Acta (BBA)-General Subjects 1425(3), 527-533.

Pizzagalli, F., Hagenbuch, B., Stieger, B., Klenk, U., Folkers, G., and Meier, P. (2002). Identification of a novel human organic anion transporting polypeptide as a high affinity thyroxine transporter. Molecular Endocrinology 16(10), 2283-2296.

Popovic, M., Zaja, R., Fent, K., and Smital, T. (2013). Molecular Characterization of Zebrafish Oatp1d1 (Slco1d1), a Novel Organic Anion-transporting Polypeptide. Journal of Biological Chemistry 288(47), 33894-33911.

Popovic, M., Zaja, R., Fent, K., and Smital, T. (2014). Interaction of environmental contaminants with zebrafish organic anion transporting polypeptide, Oatp1d1 (Slco1d1). Toxicology and applied pharmacology.

Popovic, M., Žaja, R., and Smital, T. (2010). Organic anion transporting polypeptides (OATP) in zebrafish (Danio rerio): phylogenetic analysis and tissue distribution. Comparative biochemistry and physiology. Part A, Molecular & integrative physiology 155(3), 327-335.

Poste, A. E., Hecky, R. E., and Guildford, S. J. (2011). Evaluating microcystin exposure risk through fish consumption. Environmental science & technology 45(13), 5806-5811.

Prieto, A., Jos, A., Pichardo, S., Moreno, I., and Camean, A. (2006). Differential oxidative stress responses to microcystins LR and RR in intraperitoneally exposed tilapia fish (Oreochromis sp.). Aquatic Toxicology.

Puddick, J., Prinsep, M. R., Wood, S. A., Kaufononga, S. A., Cary, S. C., and Hamilton, D. P. (2014). High Levels of Structural Diversity Observed in Microcystins from Microcystis CAWBG11 and Characterization of Six New Microcystin Congeners. Marine drugs 12(11), 5372-5395.

Quintana, N., Van der Kooy, F., Van de Rhee, M., Voshol, G., and Verpoorte, R. (2011). Renewable energy from Cyanobacteria: energy production optimization by metabolic pathway engineering. Applied microbiology and biotechnology 91(3), 471-490.

Råbergh, C., Bylund, G., and Eriksson, J. (1991). Histopathological effects of microcystin-LR, a cyclic peptide toxin from the cyanobacterium (blue-green alga) Microcystis aeruginosa on common carp (Cyprinus carpio L.). Aquatic Toxicology 20(3), 131-145.

Rabergh, C., Ziegler, K., Isomaa, B., Lipsky, M., and Eriksson, J. (1994). Uptake of taurocholic acid and cholic acid in isolated hepatocytes from rainbow trout. American Journal of Physiology-Gastrointestinal and Liver Physiology 267(3), G380-G386.

Rapala, J., Berg, K. A., Lyra, C., Niemi, R. M., Manz, W., Suomalainen, S., Paulin, L., and Lahti, K. (2005). Paucibacter toxinivorans gen. nov., sp. nov., a bacterium that degrades cyclic cyanobacterial

Page 159: Molecular cloning, characterization and relevance for

REFERENCES

- 168 -

hepatotoxins microcystins and nodularin. International Journal of Systematic and Evolutionary Microbiology 55(4), 1563-1568.

Robinson, N. A., Pace, J. G., Matson, C. F., Miura, G. A., and Lawrence, W. B. (1991). Tissue distribution, excretion and hepatic biotransformation of microcystin-LR in mice. Journal of Pharmacology and Experimental Therapeutics 256(1), 176-182.

Rodger, H., Turnbull, T., Edwards, C., and Codd, G. (1994). Cyanobacterial (blue-green algal) bloom associated pathology in brown trout, Salmo trutta L., in Loch Leven, Scotland. Journal of Fish Diseases 17(2), 177-181.

Roth, M., Obaidat, A., and Hagenbuch, B. (2012). OATPs, OATs and OCTs: the organic anion and cation transporters of the SLCO and SLC22A gene superfamilies. British journal of pharmacology 165(5), 1260-1287.

Runnegar, M., Berndt, N., and Kaplowitz, N. (1995). Microcystin uptake and inhibition of protein phosphatases: effects of chemoprotectants and self-inhibition in relation to known hepatic transporters. Toxicology and applied pharmacology 134(2), 264-272.

Runnegar, M., Seward, D. J., Ballatori, N., Crawford, J. M., and Boyer, J. L. (1999). Hepatic Toxicity and Persistence of ser/thr Protein Phosphatase Inhibition by Microcystin in the Little Skate Raja erinacea. Toxicology and applied pharmacology 161(1), 40-49.

Sachidanandam, R., Weissman, D., Schmidt, S. C., Kakol, J. M., Stein, L. D., Marth, G., Sherry, S., Mullikin, J. C., Mortimore, B. J., and Willey, D. L. (2001). A map of human genome sequence variation containing 1.42 million single nucleotide polymorphisms. nature 409(6822), 928-933.

Sahin, A., Tencalla, F. G., Dietrich, D. R., and Naegeli, H. (1996). Biliary excretion of biochemically active cyanobacteria (blue-green algae) hepatotoxins in fish. Toxicology 106(1), 123-130.

Satlin, L. M., Amin, V., and Wolkoff, A. W. (1997). Organic anion transporting polypeptide mediates organic anion/HCO3− exchange. Journal of Biological Chemistry 272(42), 26340-26345.

Schmidt, J. R., Shaskus, M., Estenik, J. F., Oesch, C., Khidekel, R., and Boyer, G. L. (2013). Variations in the Microcystin Content of Different Fish Species Collected from a Eutrophic Lake. Toxins 5(5), 992-1009.

Schopf, J., and Packer, B. (1987). Early Archean (3.3-billion to 3.5-billion-year-old) microfossils from Warrawoona Group, Australia. Science 237(4810), 70-73.

Sergeev, V., Gerasimenko, L., and Zavarzin, G. (2002). The Proterozoic history and present state of cyanobacteria. Microbiology 71(6), 623-637.

Shirasaka, Y., Mori, T., Shichiri, M., Nakanishi, T., and Tamai, I. (2012). Functional pleiotropy of organic anion transporting polypeptide OATP2B1 due to multiple binding sites. Drug metabolism and pharmacokinetics 27(3), 360-364.

Singh, S., and Asthana, R. K. (2014). Assessment of Microcystin Concentration in Carp and Catfish: A Case Study from Lakshmikund Pond, Varanasi, India. Bulletin of environmental contamination and toxicology 92(6), 687-692.

Singh, S., Kate, B. N., and Banerjee, U. (2005). Bioactive compounds from cyanobacteria and microalgae: an overview. Critical reviews in biotechnology 25(3), 73-95.

Page 160: Molecular cloning, characterization and relevance for

REFERENCES

- 169 -

Sipiä, V. O., Kankaanpää, H. T., Flinkman, J., Lahti, K., and Meriluoto, J. A. (2001). Time-dependent accumulation of cyanobacterial hepatotoxins in flounders (Platichthys flesus) and Mussels (Mytilus edulis) from the Northern Baltic Sea. Environmental toxicology 16(4), 330-336.

Sivonen, K., and Jones, G. (1999). Cyanobacterial toxins. In Toxic cyanobacteria in water: A guide to their public health consequences, monitoring and management (I. Chorus, and J. Bartram),chapter 3, pp. 26-68. Spon Press.

Smith, J. L., Boyer, G. L., and Zimba, P. V. (2008). A review of cyanobacterial odorous and bioactive metabolites: Impacts and management alternatives in aquaculture. Aquaculture 280(1), 5-20.

Smith, J. L., and Haney, J. F. (2006). Foodweb transfer, accumulation, and depuration of microcystins, a cyanobacterial toxin, in pumpkinseed sunfish (Lepomis gibbosus). Toxicon 48(5), 580-589.

Soares, R. M., Yuan, M., Servaites, J. C., Delgado, A., Magalhães, V. F., Hilborn, E. D., Carmichael, W. W., and Azevedo, S. M. (2006). Sublethal exposure from microcystins to renal insufficiency patients in Rio de Janeiro, Brazil. Environmental toxicology 21(2), 95-103.

Song, L., Chen, W., Peng, L., Wan, N., Gan, N., and Zhang, X. (2007). Distribution and bioaccumulation of microcystins in water columns: A systematic investigation into the environmental fate and the risks associated with microcystins in Meiliang Bay, Lake Taihu. Water research 41(13), 2853-2864.

Sotton, B., Devaux, A., Givaudan, N., Guillard, J., Domaizon, I., Bony, S., and Anneville, O. (2012). Short-term uptake of microcystin-LR by Coregonus lavaretus: GST activity and genotoxicity. Ecotoxicology 21(7).

Steiner, K., Hagenbuch, B., and Dietrich, D. R. (2014). Molecular cloning and functional characterization of a rainbow trout liver Oatp. Toxicology and applied pharmacology 280(3), 534-542.

Stewart, I., Seawright, A., and Shaw, G. (2008). Cyanobacterial poisoning in livestock, wild mammals and birds – an overview. In Cyanobacterial Harmful Algal Blooms: State of the Science and Research Needs (H. K. Hudnell), Vol. 619,chapter 28, pp. 613-637. Springer New York.

Stoner, R. D., Adams, W. H., Slatkin, D. N., and Siegelman, H. W. (1989). The effects of single L-amino acid substitutions on the lethal potencies of the microcystins. Toxicon 27(7), 825-828.

Sugiyama, D., Kusuhara, H., Shitara, Y., Abe, T., and Sugiyama, Y. (2002). Effect of 17β-estradiol-D-17β-glucuronide on the rat organic anion transporting polypeptide 2-mediated transport differs depending on substrates. Drug metabolism and disposition 30(2), 220-223.

Tamai, I., Nozawa, T., Koshida, M., Nezu, J.-i., Sai, Y., and Tsuji, A. (2001). Functional characterization of human organic anion transporting polypeptide B (OATP-B) in comparison with liver-specific OATP-C. Pharmaceutical research 18(9), 1262-1269.

Tencalla, F., and Dietrich, D. (1997). Biochemical characterization of microcystin toxicity in rainbow trout (Oncorhynchus mykiss). Toxicon: official journal of the International Society on Toxinology 35(4), 583.

Tencalla, F. G. (1995). Toxicity of cyanobacterial peptide toxins to fish. PhD, Swiss Federal Institute of Technology of Zürich

Tencalla, F. G., Dietrich, D. R., and Schlatter, C. (1994). Toxicity of Microcystis aeruginosa peptide toxin to yearling rainbow trout (Oncorhynchus mykiss). Aquatic Toxicology 30(3), 215-224.

Page 161: Molecular cloning, characterization and relevance for

REFERENCES

- 170 -

Tirona, R. G., Leake, B. F., Wolkoff, A. W., and Kim, R. B. (2003). Human organic anion transporting polypeptide-C (SLC21A6) is a major determinant of rifampin-mediated pregnane X receptor activation. Journal of Pharmacology and Experimental Therapeutics 304(1), 223-228.

Toranzo, A., Nieto, F., and Barja, J. (1990). Mortality associated with cyanobacterial bloom in farmed rainbow trout in Galicia (Northwestern, Spain). Bulletin- European Association of Fish Pathologists 10(4), 106-107.

Towbin, H., Staehelin, T., and Gordon, J. (1979). Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications. Proceedings of the National Academy of Sciences 76(9), 4350-4354.

Tsuji, K., Naito, S., Kondo, F., Ishikawa, N., Watanabe, M. F., Suzuki, M., and Harada, K.-i. (1994). Stability of microcystins from cyanobacteria: effect of light on decomposition and isomerization. Environmental science & technology 28(1), 173-177.

Tweedie, D., Polli, J. W., Berglund, E. G., Huang, S. M., Zhang, L., Poirier, A., Chu, X., and Feng, B. (2013). Transporter studies in drug development: experience to date and follow-up on decision trees from the international transporter consortium. Clinical Pharmacology & Therapeutics 94(1), 113-125.

van Apeldoorn, M. E., Van Egmond, H. P., Speijers, G. J., and Bakker, G. J. (2007). Toxins of cyanobacteria. Molecular nutrition & food research 51(1), 7-60.

van den Hoek, C., Mann, D., and Jahns, H. M. (1995). Algae: an introduction to phycology. Cambridge university press.

Vavricka, S. R., Van Montfoort, J., Ha, H. R., Meier, P. J., and Fattinger, K. (2002). Interactions of rifamycin SV and rifampicin with organic anion uptake systems of human liver. Hepatology 36(1), 164-172.

Vincent, W. F. (2000). Cyanobacterial dominance in the polar regions. In The ecology of cyanobacteria: their diversity in time and space (B. A. Whitton, and M. Potts),chapter 12, pp. 321-340. Springer.

Wang, L., Liang, X., Liao, W., Lei, L., and Han, B. (2006). Structural and functional characterization of microcystin detoxification-related liver genes in a phytoplanktivorous fish, Nile tilapia (Oreochromis niloticus). Comparative biochemistry and physiology. Toxicology & pharmacology: CBP 144(3), 216-227.

Wang, P., Hata, S., Xiao, Y., Murray, J. W., and Wolkoff, A. W. (2008). Topological assessment of oatp1a1: a 12-transmembrane domain integral membrane protein with three N-linked carbohydrate chains. American Journal of Physiology-Gastrointestinal and Liver Physiology 294(4), G1052-G1059.

Ward, D. M., and Castenholz, R. W. (2000). Cyanobacteria in geothermal habitats. In The ecology of cyanobacteria: their diversity in time and space (B. A. Whitton, and M. Potts),chapter 3, pp. 37-59. Springer.

Watanabe, M., Oishi, S., Harda, K., Matsuura, K., Kawai, H., and Suzuki, M. (1987). Toxins contained in Microcystis species of cyanobacteria (blue-green algae). Toxicon: official journal of the International Society on Toxinology 26(11), 1017-1025.

Weckesser, J., Drews, G., and Mayer, H. (1979). Lipopolysaccharides of photosynthetic prokaryotes. Annual Reviews in Microbiology 33(1), 215-239.

Whitton, B. A. (2000). Soils and rice-fields. In The ecology of cyanobacteria: their diversity in time and space (B. A. Whitton, and M. Potts),chapter 8, pp. 233-255. Springer.

Page 162: Molecular cloning, characterization and relevance for

REFERENCES

- 171 -

World Health Organization (1998). Guidelines for Drinking-Water Quality. 2. 2. Health Criteria and Other Supporting Information.

Wickstrom, M. L., Khan, S. A., Haschek, W. M., Wyman, J. F., Eriksson, J. E., Schaeffer, D. J., and Beasley, V. R. (1995). Alterations in microtubules, intermediate filaments, and microfilaments induced by microcystin-LR in cultured cells. Toxicologic pathology 23(3), 326-337.

Williams, D., Craig, M., Dawe, S., Kent, M., Andersen, R., and Holmes, C. (1997). 14C-labeled microcystin-LR administered to atlantic salmon via intraperitoneal injection provides in vivo evidence for covalent binding of microcystin-LR in salmon livers. Toxicon 35(6), 985-989.

Williams, D. E., Kent, M. L., Andersen, R. J., Klix, H., and Holmes, C. F. (1995). Tissue distribution and clearance of tritium-labeled dihydromicrocystin-LR epimers administered to Atlantic salmon via intraperitoneal injection. Toxicon 33(2), 125-131.

Xie, L., Xie, P., Guo, L., Li, L., Miyabara, Y., and Park, H. D. (2005). Organ distribution and bioaccumulation of microcystins in freshwater fish at different trophic levels from the eutrophic Lake Chaohu, China. Environmental toxicology 20(3), 293-300.

Xu, L., Lam, P., Chen, J., Zhang, Y., and Harada, K. (2000). Comparative study on in vitro inhibition of grass carp (Ctenopharyngodon idellus) and mouse protein phosphatases by microcystins. Environmental toxicology 15(2), 71-75.

Xue, Y., Ren, J., Gao, X., Jin, C., Wen, L., and Yao, X. (2008). GPS 2.0, a tool to predict kinase-specific phosphorylation sites in hierarchy. Molecular & Cellular Proteomics 7(9), 1598-1608.

Yamaguchi, H., Sugie, M., Okada, M., Mikkaichi, T., Toyohara, T., Abe, T., Goto, J., Hishinuma, T., Shimada, M., and Mano, N. (2010). Transport of estrone 3-sulfate mediated by organic anion transporter OATP4C1: estrone 3-sulfate binds to the different recognition site for digoxin in OATP4C1. Drug metabolism and pharmacokinetics 25(3), 314-317.

Yan, W., Zhou, Y., Yang, J., Li, S., Hu, D., Wang, J., Chen, J., and Li, G. (2012). Waterborne exposure to microcystin-LR alters thyroid hormone levels and gene transcription in the hypothalamic–pituitary–thyroid axis in zebrafish larvae. Chemosphere 87(11), 1301-1307.

Yoshizawa, S., Matsushima, R., Watanabe, M. F., Harada, K.-i., Ichihara, A., Carmichael, W. W., and Fujiki, H. (1990). Inhibition of protein phosphatases by microcystis and nodularin associated with hepatotoxicity. Journal of cancer research and clinical oncology 116(6), 609-614.

You, G., and Morris, M. E. (2007). Drug Transporter: Molecular Characterization and Role in Drug Disposition. Wiley.

Zeck, A., Weller, M. G., Bursill, D., and Niessner, R. (2001). Generic microcystin immunoassay based on monoclonal antibodies against Adda. Analyst 126(11), 2002-2007.

Zhang, D., Ping, X., Chen, J., Dai, M., Qiu, T., Liu, Y., and Liang, G. (2009a). Determination of microcystin-LR and its metabolites in snail (Bellamya aeruginosa), shrimp (Macrobrachium nipponensis) and silver carp (Hypophthalmichthys molitrix) from Lake Taihu, China. Chemosphere 76(7), 974-981.

Zhang, D., Xie, P., Liu, Y., and Qiu, T. (2009b). Transfer, distribution and bioaccumulation of microcystins in the aquatic food web in Lake Taihu, China, with potential risks to human health. Science of the Total Environment 407(7), 2191-2199.

Page 163: Molecular cloning, characterization and relevance for

REFERENCES

- 172 -

Zhang, D., Yang, Q., Xie, P., Deng, X., Chen, J., and Dai, M. (2012). The role of cysteine conjugation in the detoxification of microcystin-LR in liver of bighead carp (Aristichthys nobilis): a field and laboratory study. Ecotoxicology 21(1), 244-252.

Zhang, W., Liang, G., Wu, L., Tuo, X., Wang, W., Chen, J., and Xie, P. (2013). Why mammals more susceptible to the hepatotoxic microcystins than fish: evidences from plasma and albumin protein binding through equilibrium dialysis. Ecotoxicology 22(6), 1012-1019.

Zhang, X., Ji, W., Zhang, H., Zhang, W., and Xie, P. (2011). Studies on the toxic effects of microcystin-LR on the zebrafish (Danio rerio) under different temperatures. Journal of Applied Toxicology 31(6), 561-567.

Zhao, S., Xie, P., Li, G., Jun, C., Cai, Y., Xiong, Q., and Zhao, Y. (2012). The proteomic study on cellular responses of the testes of zebrafish (Danio rerio) exposed to microcystin-RR. Proteomics 12(2), 300-312.

Zimba, P., Khoo, L., Gaunt, P., Brittain, S., and Carmichael, W. (2001). Confirmation of catfish, Ictalurus punctatus (Rafinesque), mortality from Microcystis toxins. Journal of Fish Diseases 24(1), 41-47.

Zimba, P. V., Camus, A., Allen, E. H., and Burkholder, J. M. (2006). Co-occurrence of white shrimp,Litopenaeus vannamei, mortalities and microcystin toxin in a southeastern USA shrimp facility. Aquaculture 261(3), 1048-1055.