108
Louisiana State University LSU Digital Commons LSU Doctoral Dissertations Graduate School 2012 Osmoregulation and acid-base tolerance in fish Shujun Zhang Louisiana State University and Agricultural and Mechanical College, [email protected] Follow this and additional works at: hps://digitalcommons.lsu.edu/gradschool_dissertations is Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please contact[email protected]. Recommended Citation Zhang, Shujun, "Osmoregulation and acid-base tolerance in fish" (2012). LSU Doctoral Dissertations. 3362. hps://digitalcommons.lsu.edu/gradschool_dissertations/3362

Osmoregulation and acid-base tolerance in fish

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Osmoregulation and acid-base tolerance in fish

Louisiana State UniversityLSU Digital Commons

LSU Doctoral Dissertations Graduate School

2012

Osmoregulation and acid-base tolerance in fishShujun ZhangLouisiana State University and Agricultural and Mechanical College, [email protected]

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_dissertations

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion inLSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please [email protected].

Recommended CitationZhang, Shujun, "Osmoregulation and acid-base tolerance in fish" (2012). LSU Doctoral Dissertations. 3362.https://digitalcommons.lsu.edu/gradschool_dissertations/3362

Page 2: Osmoregulation and acid-base tolerance in fish

I

OSMOREGULATION AND ACID-BASE TOLERANCE IN FISH

A Dissertation

Submitted to the Graduate Faculty of the

Louisiana State University and

Agricultural and Mechanical College

in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

in

The Department of Biological Sciences

by

Shujun Zhang

B.S., Huazhong University of Science & Technology, 2003

M.S., Huazhong Unversity of Science & Technology, 2005

August 2012

Page 3: Osmoregulation and acid-base tolerance in fish

II

ACKNOWLEDGMENTS

I would like to thank my major professor, Dr. Fernando Galvez, for his

devotion of time on mentoring my research, financial support, and his guidance

on the way to enjoy research. I appreciate his sincere and warm-hearted help

which makes my life much happier.

I would like to thank my committee members, Dr. Andrew Whitehead, Dr.

Steve C. Hand, Dr. Evanna Gleason, Dr. Gentry Glen T for their approval of my

research project, sharing their laboratory facilities, their advice, and their

stimulating discussion in my research.

I would like to thank Dr. Charlotte M Bodinier, Benjamin Dubansky, Ying

Guan, Yanling Meng, Christine Savolainen, Charles Brown, and my friends in the

Department of Biological Sciences. Their friendship and help in the experiments

made my time in the lab much easier.

My special thanks to my brother and sister, Zheng and Jing, for their never-

ending support, no matter if I am in high or low.

Lastly, my sincere thanks go to my parents, Dong and Feng, for their love

and devotion during my study; without their support I would not have finished my

Ph.D. Research.

Page 4: Osmoregulation and acid-base tolerance in fish

III

TABLE OF CONTENTS

ACKNOWLEDGMENTS……………………...………………………………………II

LIST OF TABLES…………………………………………………….………………....V

LIST OF FIGURES……………………………………………………………………VI

ABSTRACT…………………………………………………………….…………….VIII

CHAPTER 1 LITERATURE REVIEW

General Principles of Osmoregulation in Teleost Fish……………....1

Ion Transport Models in the Gills of Freshwater (FW)-Acclimated Fish……………………………………………………............................3

Ion Transport Models in the Gills of Seawater (SW)-Acclimated Fish…………………………………………………………………..…...10

Fundulus as the Premier Model in Environmental Biology and Fish Physiology……………………………………………...........................12

Significance of Paracellular Pathways in Osmoregulation ……......15

General Principles of Acid-base Tolerance in Teleost Fish………..18

Brief Outline of Main Points of Each Chapter……….………….….20

CHAPTER 2 ALTERATIONS IN CLAUDIN EXPRESSIONS IN THE GILLS OF FUNDULUS GRANDIS FOLLOWING ABRUPT SALINITY TRANSFER

Introduction…………………..……...………………………………..……23

Materials and Methods……..……...……………………….………..……26

Results………. .. ………………………………...……….…….……..…..38

Discussion……..……………………………….……….…..…….…..……48

CHAPTER 3 ACID-BASE TOLERANCES AMONG FUNDULUS SPECIES

Introduction………………………………………..…………………....…59

Materials and Methods……..……...……….………………………....…62

Results………. .. …………………………………...………………..…..68

Discussion……..………………….……………….……..……….………73

Page 5: Osmoregulation and acid-base tolerance in fish

98……………….

3

0

IV

CHAPTER 4 PROSPECTIVES………………………………………..……..8

REFERENCES CITED………………………………………………………..…....8

VITA………………………………………………………………..…

Page 6: Osmoregulation and acid-base tolerance in fish

V

LIST OF TABLES

Table 2.1 Ion concentration (µM) in the waters used in the experiments………28

Table 2.2 Sequences of primers used for quantitative PCR………………………37

Page 7: Osmoregulation and acid-base tolerance in fish

VI

LIST OF FIGURES

Figure 1.1: Traditional model for NaCl uptake by FW-adapted teleost fish gi l ls………………………………………… .……………………………………4

Figure 1.2: Traditional model for NaCl secretion in SW-acclimated fish gills……………………………………………………………… …………….10

Figure 1.3: A typical model showing the structure of tight junction proteins in paracellular pathways between two adjacent cells ……………………17

Figure 2.1: Plasma sodium (A) (N=6) and chloride (B) (N=6) levels in Fundulus g r a n d i s … … … … … … … … … … … … . … … … … … … … … . … … … 3 9

Figure 2.2: Total (net) flux and unidirectional fluxes of Na+ (A) and Cl- (B) in Fundulus grandis after transfer from 5 ppt water to 0.1 ppt water (N=6) versus t i m e . . … … … … … … … … … … … … … … … … … … … … … … … . . … … 4 0

Figure 2.3: Gill SEM images of Fundulus grandis.. …………......…….….42

Figure 2.4: Total g i l l permeabi l i ty rate of PEG-4000 in Fundulus grandis……………………………………………………………………….…43

Figure 2.5: Junctions between MRCs and adjacent cells in Fundulus grandis 3 days af ter the t ransfer f rom 5 ppt water to 5 , 2, 1, 0 .5, 0.1 ppt water………………………………………………………………….. ………44

Figure 2.6: The effects of hypoosmotic exposure (0.1 ppt and 0.5 ppt exposure) on the abundance of Cldn3, Cldn5, Cldn23, Cldn28, Cldn7 and Cldn26 in gills of F u n d u l u s g r a n d i s … . . … … … … … … … … … … … . . . … … . 4 5

Figure 2.7: The effects of hypoosmotic exposure (0.1 ppt exposure) on the abundance of Cldn3, Cldn5, Cldn23, Cldn28, Cldn7 and Cldn26 in two different ep i the l ia l ce l l s : PVCs and MRCs . . …………………… . .…….…. .47

Figure 3.1: Na+ influx, efflux and net flux rates in Fundulus heteroclitus MDPP at 0 h, 12 h, 24 h, 3 day and 7 day point during a 7-day acid challenge p e r i o d … … … … … … … … … … … … … … … … … … … … … . . … … . . 6 9

Figure 3.2: Na+ influx, efflux and net flux rates in Fundulus heteroclitus VAcoast at 0 h, 12 h, 24 h, 3 day and 7 day point during a 7-day acid challenge p e r i o d … … … … … … … … … … … … … … … … … … … … … … … . . . 7 0

Page 8: Osmoregulation and acid-base tolerance in fish

VII

Figure 3.3: Na+ influx, efflux and net flux rates in Fundulus majalis after HCl i n j e c t i o n … … … … … … … … … … … … … … … … … . . … … … … … … … . 7 2

Figure 3.4: Claudin 3, -5, -23, -28 mRNA expression level changes in Fundulus h e t e r o c l i t u s M D P P g i l l s d u r i n g a 7 - d a y a c i d c h a l l e n g e per iod………………… ………………………………………………….. …73

Figure 3.5: Claudin 3, -5, -23, -28 mRNA expression level changes in Fundulus h e t e r o c l i t u s V A c o a s t g i l l s d u r i n g a 7 - d a y a c i d c h a l l e n g e per iod……………………………………………………………………. ….74

Page 9: Osmoregulation and acid-base tolerance in fish

VIII

ABSTRACT

The gill, with a large surface area in intimate contact with the environment,

is the primary organ of ion and acid-base regulation in fish. These same

characteristics make the gill epithelium particularly susceptible to a number of

environmental perturbations, including osmotic challenges and metabolic

acidosis. For most freshwater fish, the active uptake of the strong ions, sodium

and chloride, are intimately linked with the excretion of acid and base equivalents,

respectively. However, fish from the genus Fundulus, are unique in their

apparent lack of active chloride uptake at the gills. This unique feature makes

limiting Cl- loss through paracellular pathways the only practical strategy for

these species in tolerating acute exposure to hypoosmotic conditions especially

when dietary chloride is limited. We find that Fundulus grandis can dynamically

regulate their paracellular pathway in the gill epithelium at salinities approaching

fresh water. We observe the significant up-regulation in the mRNA levels of

several claudins in the gill and that some of these claudins exhibit expressional

discrepancies between mitochondrion-rich and pavement cells in fish gills, which

shows correlations with changes in gill morphology and ion flux rates in fish.

Collectively, our data suggest that claudins may play important roles in regulating

ion flux across paracellular pathways of Fundulus during osmotic challenges.

The linkage between osmoregulation and acid-base tolerance has long

been studied in teleost fish. Failure to identify an active Cl- uptake system in

freshwater Fundulus indicates the uniqueness of these species in

osmoregulation and acid-base tolerance. We find that Fundulus heteroclitus

Page 10: Osmoregulation and acid-base tolerance in fish

IX

increases Na+ uptake within the first few hours of metabolic acidosis, We also

find that metabolic acidosis induces the changes of mRNA levels of several gill

claudin proteins, which may play roles in regulating the permeability of the

paracellular pathway to strong ions. Our data show Fundulus heteroclitus is

capable to cope with great metabolic acidosis within their bodies, which may

contribute to their adaptation to internal and external perturbations.

Page 11: Osmoregulation and acid-base tolerance in fish

1

CHAPTER 1: LITERATURE REVIEW

General Principles of Osmoregulation in Teleost Fish

Fish are the largest and most diverse group of vertebrates, with an

estimated 20,000 species inhabiting varied aquatic environments worldwide

(Vernier, 1989). The majority of fish are stenohaline based on their ability to

tolerate only relatively narrow salinity fluctuations. In contrast, a comparably

smaller proportion of fish species are referred to as euryhaline, due to their

capacity to tolerate large extremes in environmental salinity (Krogh, 1939). Even

among euryhaline fishes, there is large diversity amongst fish in their capacity to

tolerate salinity extremes (Hwang and Lee, 2007). Some euryhaline species,

such as salmonids, are capable of making these salinity transitions only over

extended periods of time. In the case of salmonids, they begin their lives in fresh

water (FW), move from FW to sea water (SW) after smoltification, and then move

back to FW often after a few years (McCormick et al., 1991). Other species,

which are typically found in intertidal regions, tolerate more frequent salinity

transfers. A classic example includes fish from the genus Fundulus from the

family Fundulidae (Burnett et al., 2007a).

The physiology of euryhaline fish changes dramatically depending on

whether the fish is located in FW, which is hypoosmotic to the extracellular fluids

of the animal, or in SW, which is hyperosmotic to its internal environment (Evans

and Claiborne, 2006; Evans et al., 1999; Hoffmann, 1992). In the former case,

fish need to absorb ions actively from the external environment to compensate

for the passive loss of ions to FW. In the latter case, animals actively excrete ions

to offset the passive loading of ions from SW. In contrast to stenohaline fishes,

Page 12: Osmoregulation and acid-base tolerance in fish

2

many euryhaline species are able to tolerate dramatic shifts in environmental

salinity by making physiological adjustments over the course of hours to days to

facilitate restoration of ion homeostasis (Evans et al., 2005).

Many studies have focused on euryhaline fishes in order to elucidate the

osmoregulatory mechanisms that allow them to cope with fluctuations in

environmental salinity(Evans, 2006; Evans, 2008; Evans and Claiborne, 2006;

Hwang and Lee, 2007; Wood and Marshall, 1994). It is generally accepted that

the gill epithelium is the primary osmoregulatory organ in fish, with the

gastrointestinal tract and kidneys also playing important supporting roles, which

depend on environmental salinity (Evans et al., 2005). FW-acclimated fish

actively absorb ions from the external environment at their gills and minimize ion

loss by reducing the paracellular movement of ions at their body surfaces, while

actively absorbing ions at their kidney to minimize urinary ion loss. SW-

acclimated fish drink SW, from which they actively absorb ions to facilitate water

absorption in the intestine. This salt load is subsequently secreted actively at the

gills, utilizing a process of transcellular chloride transport and paracellular sodium

efflux (Evans, 2006; Evans, 2008; Evans and Claiborne, 2006) . Regardless of

environmental salinity, the gills are extremely important for osmotic regulation in

euryhaline fish (Burnett et al., 2007a; Evans, 2008).

The fish gills are composed of a variety of different cell types, although the

mitochondrion-rich (MR) cells and the pavement (PV) cells are the ones most

often implicated in osmoregulatory functions. In particular, the MR cells are

critical in the active extrusion of ions in SW-acclimated animals, and are

Page 13: Osmoregulation and acid-base tolerance in fish

3

generally associated with active ion uptake in FW fish (Chang et al., 2001; Evans

et al., 2005), although their phenotypes vary extensively with salinity. In FW, MR

cells have large surface areas with small villi protruding to external environments.

In contrast, in SW, MR cells hide underneath the gill epithelial surface in

connection to external environments through a small crypt(Wilson and Laurent,

2002). FW- and SW-acclimated fishes possess a wide variety of ion transport

proteins, which mediate the active and passive movement of ions. MR cells in

both FW- and SW-acclimated fish express many of the same ion transporters,

although these proteins differ in their cellular localization (i.e., apical versus

basolateral membrane expression), and in their regulation (Hwang and Lee,

2007).

Ion Transport Models in the Gills of Freshwater (FW)-Acclimated Fish

Ever since the late1930‟s, there has been general agreement that the

unidirectional absorptions of Na+ and Cl- in the freshwater fish gill are

independent of one another and linked to the extrusion of proton and bicarbonate,

respectively (Krogh, 1938, 1939; Parry et al., 1959). However, there is still

considerable debate on the cellular localization and molecular underpinnings of

these transport processes (Hwang and Lee, 2007). Although an exhaustive

review of these transport models is beyond the scope of my dissertation, I would

like to highlight some of the major details of the most widely accepted models

(Fig 1.1), and describe the current understanding associated with the role of

these transports in the fish gill of FW-acclimated Fundulus genus, which is the

species of focus for my dissertation.

Page 14: Osmoregulation and acid-base tolerance in fish

4

Figure 1.1: Traditional model for NaCl uptake by FW-adapted teleost fish gills. The distribution of transport proteins in specific cells may vary with fish species. (Figure modified from (Evans, 2010)).

Ion transport in the gills of most FW teleost exhibit the inextricable link of

Na+ uptake and H+ excretion either through a Na+/H+ exchanger (NHE) or

through a putative apical Na+ channel, electrochemically linked to a V-type H+-

ATPase on the fish gill (Edwards and Toop, 2002; Hwang and Lee, 2007;

Inokuchi et al., 2008). NHE transports Na+ from the water into the cell in

exchange for proton (or NH4+). Other studies have suggested that un-ionized NH3

combines with proton to produce NH4+, which is trapped after leaving the cell,

thus preventing the back flux of H+ (Hwang and Lee, 2007; Perry et al., 2003a;

Perry et al., 2003b; Tresguerres et al., 2005). Either way, this Na+ and proton

exchange (either H+ directly or H+ that is chemically combined with NH3) is

electro-neutral, relying solely on the prevailing electrochemical gradients for both

Na+ entry and proton excretion. At first glance, these mechanisms of exchange

Page 15: Osmoregulation and acid-base tolerance in fish

5

appear problematic. The low concentrations of Na+ (e.g., low µM Na+

concentrations of some fresh waters) would in fact represent a negative driving

force for Na+ entry (George et al., 2006b). Furthermore, the intracellular proton

concentrations of fish gill cells (pH 7.4) would be insufficient to stimulate Na+

uptake, especially under scenarios of environmental acidification (Evans, 2008;

Goss and Wood, 1990a).

An alternate model of transepithelial Na+ uptake in the FW fish gill was

proposed involves the indirect coupling of Na+ and H+ exchange at the apical

membrane (Avella and Bornancin, 1989). This mechanism, which was found to

exist in the turtle bladder and skin of FW-acclimated frogs (Duranti et al., 1986),

had Na+ permeating the apical membrane via an epithelial Na+ channel (ENaC)

(Evans, 2008) energized by an apical V-type H+-ATPase. The active extrusion of

H+ via a V-type H+-ATPase would create a localized negative charge along the

apical membrane fueling passive Na+ entry against a concentration gradient.

Intracellular Na+ would then exit the cytoplasm into plasma via Na+/K+ ATPase

(NKA). This mechanism appeared to alleviate concerns raised for the NHE model,

which related to the absence of physiologically-relevant electrochemical

gradients. Several physiological, radio-isotopic, pharmacological,

immunocytochemical, cellular and molecular techniques have provided

compelling evidence of the existence of a V-type H+-ATPase in the fish gill

(Evans et al., 2005; Hwang and Lee, 2007). For example, bafilomycin, a specific

V-type H+-ATPase inhibitor inhibits Na+ uptake by up to 90% in tilapia and

zebrafish (Biosen et al., 2003; Fenwick et al., 1999). Immunocytochemical

Page 16: Osmoregulation and acid-base tolerance in fish

6

studies using heterologous antibodies raised against various subunits of V-type

H+-ATPase have localized the protein to the gills of rainbow trout, zebrafish,

shark, and tilapia (Biosen et al., 2003; Sullivan et al., 1995; Wilson et al., 2000;

Wilson et al., 2002). Vanadate, an inhibitor of the P-type H+-ATPase inhibits

proton secretion by approximately 50% in rainbow trout, which further indicates

the existence of a V-type H+-ATPase (Lin and Randall, 1991). Despite all the

evidence supporting this model, the existence of ENaC is still in question based

on the inability to identify a genomic sequence for ENaC in any of the fully-

sequenced genomes of teleost fish. In fact, in Fundulus heteroclitus, V-type H+-

ATPase localizes to the basolateral membrane of the fish gill, where it is thought

to play a small role in Cl- absorption in FW (Fumi et al., 2002). In this case,

carbonic anhydrase breaks down CO2 into H+ and HCO3-, and H+ is pumped into

plasma facilitating HCO3- exit from cytoplasm (Fumi et al., 2002).

While molecular techniques have put into question the universal adherence

of the fish gill to this ENaC/ V-type H+-ATPase model, resurgence in the putative

role of a NHE in transepithelial Na+ transport has once again emerged (Hwang

and Lee, 2007). One study identified NHE2 mRNA in the gill of longhorn sculpin

and killifish, and another study localized NHE3 to MR cells in rainbow trout and

blue-throated wrasse (Edwards and Toop, 2002). Amiloride, an inhibitor to NHE,

blocks Na+ accumulation by over 90% in zebrafish embryos (Esaki et al., 2007).

Though there is evidence for the existence of NHE, the underlying driving force

still remains unclear. Some propose that carbonate anhydrase (CA) plays a key

role in this process by converting metabolic CO2 into HCO3- and H+. Some

Page 17: Osmoregulation and acid-base tolerance in fish

7

propose that carbonate anhydrase (CA) plays a key role in this process by

converting metabolic CO2 into HCO3- and H+ which may be secreted to external

environments by NHE. A recent study found that the CA inhibitor, ethoxzolamide,

reduced Na+ uptake in zebrafish (Biosen et al., 2003). However, one study found

that CA is only expressed in non-MR cells in dogfish gill (though more studies

confirmed that CA localized to the apical side of both MR cells and pavement

cells), which represents a cellular distribution inconsistent with the proposed

model (Gilmour et al., 2006).

One of widely accepted models of unidirectional Cl- uptake for FW fish

involves the absorption of Cl- in exchange for HCO3- via an apical anion

exchanger (AE)(Evans, 2010; Hwang and Lee, 2007; Shmukler et al., 2005;

Tang and Lee, 2007). A Cl-/HCO3- exchanger has been immunolocalized to the

apical surface of branchial cells in tilapia, although the cellular distribution

remains somewhat uncertain. In one study, an oligonucleotide probe

(complementary to rat AE cDNA) hybridized to RNA in both the filament and

lamellae of the rainbow trout, suggesting that both pavement cells and MR cells

contain mRNA transcripts for Cl-/HCO3- exchanger (Gary et al., 1996). However,

tilapia anion exchanger 1 (AE1) and the basolateral NKA have been co-localized

to the gills of pufferfish of FW fish (Wilson et al., 2000; Wilson et al., 2002). As

stated previously, carbonic anhydrase (CA) has been identified in fish branchial

cells, and is also presumed to be the driving force for Cl- uptake, in that it

provides HCO3- from metabolic CO2. One study found that low waterborne Cl-

stimulates Cl- uptake and enhances CA protein expression in the gills of tilapia,

Page 18: Osmoregulation and acid-base tolerance in fish

8

suggesting a correlation between CA expression and unidirectional Cl- uptake

(Chang and Hwang, 2004).

Although Cl-/HCO3- exchanger is considered the main mechanism of Cl-

uptake in FW-adapted fish, Na+/K+/2Cl- co-transport (NKCC) and Na+/Cl- co-

transport (NCC) are also considered as candidates for the apical Cl- uptake

(Evans et al., 2005; Hwang and Lee, 2007). In a study of tilapia embryonic skin

and adult gills, a heterologous antibody revealed the presence of apical NKCC in

FW-type MR cells, and apical NKCC was suggested to be associated with

Na+/Cl- uptake (Hiroi et al., 2005). It is worthy to mention that an active Cl- uptake

system has not been identified in killifish, which is a departure from general fish

model(Burnett et al., 2007b; George et al., 2006b; Patrick et al., 1997; Wood and

Marshall, 1994).

NKA is the primary driving force for transepithelial Na+ transport in fish gill,

but is also important in energizing indirectly the active uptake of other ions

(Hwang and Lee, 2007; Perry et al., 2003b). Some studies have also suggested

that a Na+/HCO3- cotransporter (NBC) in the basolateral membrane is involved in

this step too. One study found expression of a NBC in branchial MR cells in

Osorezan dace (Hirata et al., 2003). In another study, an isoform of the NBC was

identified in the gill tissue of rainbow trout (NCBI GenBank accession ID:

AF434166, NBC protein sequence in rainbow trout) (Perry et al., 2003b). NBC

could potentially mediate both the Na+ and Cl- transport across the basolateral

membrane of the gills, which facilitates Cl- enter plasma through basolateral

membrane channels. Some studies suggest that specific Cl- channels exist (Tang

Page 19: Osmoregulation and acid-base tolerance in fish

9

and Lee, 2007). One study has demonstrated a higher protein expression of

CLC-3 Cl- channels in the gills of freshwater puffer fish than in seawater

fish(Tang and Lee, 2007). However, though CLC-3 Cl- channels have been

cloned and expressed in the gills of tilapia, they have been proposed to function

in intracellular Cl- regulation rather than in transepithelial Cl- uptake based on in

vitro functional analyses (Hiroaki et al., 2002). Finally, a cystic fibrosis

transmembrane regulator (CFTR) has also been proposed to be another

candidate for basolateral Cl- transport. Marshall et al localized the CFTR channel

in both pavement cells and MR cells in Fundulus heteroclitus, suggesting that

CFTR might be involved in Cl- exit into plasma (Marshall et al., 2002). Further

studies apparently are needed to verify these competing results.

Early reports have suggested that both Na+ and Cl- transports are

performed in a single mitochondrion-rich ionocyte, although more recent work

proposes that at least 2 or more functionally-distinct MR cell types coordinate

these active ion transports (Evans et al., 2005; Galvez et al., 2001; Goss et al.,

2001; Hwang and Lee, 2007; Katoh et al., 2001; Scott et al., 2004b).

A point needed to mention is that a single model does not apply to all the teleost

(Evans, 2008; Marshall and Grosell, 2005). Until now there is still no

characterization of a chloride uptake pathway in apical side of gills of Fundulus

species(Patrick et al., 1997; Patrick and Wood, 1999). Some studies indicated

that Fundulus species maintain plasma chloride by decreasing passive chloride

loss from their bodies (Patrick et al., 1997). This is a departure from the general

proposed model. One study suggests that Fundulus species maintained Cl-

Page 20: Osmoregulation and acid-base tolerance in fish

10

concentration by decreasing the passive loss of the ion by regulating paracellular

tight junctions (Scott et al., 2004b). This is a new strategy of maintaining Cl-

homeostasis in plasma. This dissertation will address special emphasis on

important roles of paracellular pathways regulating ion transport in

osmoregulation among Fundulus species.

Ion Transport Models in the Gills of Seawater (SW)-Acclimated Fish

Figure 1.2: Traditional model for the mechanism of salt secretion in branchial epithelia of SW-acclimated fish. The distribution of transport proteins in specific cells may vary with fish species. MRC represents mitochondrion-rich cells, CFTR represents cystic fibrosis transmembrane conductance regulators in this figure. (Figure modified from (Evans, 2010)).

SW-acclimated fish actively secrete ions at the gills to counteract the

accumulation of ions from ingested SW in the intestine (Evans et al., 2005). The

gills of SW-acclimated fish express NKA and Na+/K+/2Cl- co-transporter (NKCC)

on the basolateral membrane, and the cystic fibrosis transmembrane regulator

Cl- channel (CFTR) on the apical membrane (Figure 1.2). NKA serves as the

driving force for osmoregulation by pumping Na+ out of branchial cells and

Page 21: Osmoregulation and acid-base tolerance in fish

11

absorbing K+ into cells. This maintains a low intracellular Na+ concentration and

indirectly leads to the establishment of a negative intracellular membrane

potential, which facilitate Na+ and Cl- entry via NKCC. Although intracellular Na+

is recycled back through NKA, the resulting increase in intracellular Cl- and the

highly-negative membrane potential facilitate the passive diffusion of Cl- through

the apical CFTR. A positive, blood-side transepithelial potential creates the

electromotive force for efflux of Na+ into SW via a leaky paracellular pathway

between MR cells and accessory cells (AC) (Hwang and Lee, 2007). The

permeability of this leaky junction is partly controlled by claudins which will be

addressed later(Koval, 2006; Van Itallie and Anderson, 2006). NKA is important

for salt transport not only because of its role in Na+ transport but also because it

is the main driving force for the whole salt secretion process. NKCC is

considered as the site of Na+ and Cl- entry into cytoplasm facilitated by the

electrochemical gradient created by NKA in fish osmoregulation. High expression

of NKCC is usually correlated with seawater acclimation in fish. One study

showed NKCC mRNA levels were increased in Fundulus heteroclitus gills

following transfer from brackish water to sea water(Scott et al., 2004a). Another

study found that NKCC expression in bass gills changed in response to salinity

change(Tipsmark et al., 2004). An immunofluorescence study showed the

expression of NKCC in MR cells in Fundulus heteroclitus too(Marshall et al.,

2002).

In terms of Cl- secretion, the Cl- CFTR has been implicated in transcellular

Cl- secretion in marine teleost since its initial discovery in the gills of Fundulus

Page 22: Osmoregulation and acid-base tolerance in fish

12

heteroclitus(Singer et al., 1998). Time course studies with Fundulus heteroclitus

reveal that expression of apical CFTR mRNA in gills appears at 24h after the

transfer from FW to 10% SW while CFTR expression in gills decrease as fish

are transferred to FW(Scott et al., 2005; Tipsmark et al., 2002). CFTR mRNA

levels ingills of Hawaiian goby also increases during SW acclimation(McCormick

et al., 2003). Overall, Cl- secretion from the cytoplasm to sea water is facilitated

by the inside negative membrane potential of the gill cells.

Fundulus as the Premier Model in Environmental Biology and Fish Physiology

Fundulus genus consists of approximately 27 species in North America,

with about 11 of them distributed along the coasts of the Atlantic Ocean and the

Gulf of Mexico, or in inland regions of North America(Burnett et al., 2007a;

Griffith, 1974b; Whitehead et al., 2010).

Fundulus species (also called killifish) usually reside in brackish estuaries

and salt marshes along coastal regions, however, many of these species are

renowned for their abilities to tolerate salinity extremes ranging from FW to

approximately 3- to 4-fold times the strength of SW (Griffith, 1974b; Loeb and

Wasteneys, 1912; Sumner, 1911). Some Fundulus species are often trapped

during ebb tide in small ponds, which may become more saline due to

evaporative water loss, or more dilute due to large rainstorm events(Marshall,

2003; Marshall et al., 2000; Marshall et al., 1999; Patrick et al., 1997). As such,

they have served as an excellent model for studying osmoregulatory physiology

and stress biology (Burnett et al., 2007b). Although most Fundulus species are

thought to have regulatory mechanisms to tolerate these dramatic salinity

Page 23: Osmoregulation and acid-base tolerance in fish

13

gradients, other Fundulus exclusively live in FW or SW, but not both, and thus,

are less tolerant of fluctuations in environmental salinity (Griffith, 1974b;

Whitehead, 2010).

Of these Fundulus species, Fundulus heteroclitus is most renowned for its

environmental plasticity, thus widely used to extend the understanding of

strategies to cope with environmental challenges such as changes in salinity,

oxygen levels, and temperature(Burnett et al., 2007b; Fangue et al., 2006;

George et al., 2006a). Fundulus heteroclitus does not migrate during its life cycle,

with local sub-populations exhibiting summer home ranges on the order of 30-40

m and restricted winter movements (Burnett et al., 2007a; Chidester, 1920).

These characteristics of broad distribution and limited habitat range have made

Fundulus heteroclitus a popular and powerful field model for studying biological

and ecological plasticity to natural environmental challenges. In addition, there is

evidence of local adaptation to environmental pollutants of Fundulus heteroclitus

residing in heavily polluted waters. Consequently, this species, as well as other

Fundulus species, are used in both laboratory and field studies to elucidate some

disease processes and toxicological mechanisms, are used in both laboratory

and field studies to elucidate some disease processes and toxicological

mechanisms as well (Burnett et al., 2007a).

Fundulus species have been studied extensively in the past 60 years,

revealing many basic principles that are now accepted as fundamental to our

knowledge of euryhaline osmoregulation in teleost fish (Hossler et al., 1985;

Karnaky, 1980, 1986; Karnaky et al., 1984; Karnaky et al., 1976). Ion transport

Page 24: Osmoregulation and acid-base tolerance in fish

14

across the gills of SW-acclimated Fundulus occurs like it does in other marine

teleosts. Cystic fibrosis transmembrane conductance Cl- regulator (CFTR),

Na+/K+/2Cl- co-transporter (NKCC), and NKA coordinate the transcellular

secretion of Cl- in MR cells, setting up for the concomitant movement of Na+ into

water via the leaky-type tight junctions between MR cells and accessory cells. In

contrast, the ion transport model for FW-acclimated Fundulus is a departure from

that of most other fish species (Patrick and Wood, 1999). Generally, the active

uptake of Cl- in FW teleost is linked with HCO3- excretion in MR cells, via a

presumptive Cl-/HCO3- exchanger. This process however does not apply to FW-

acclimated Fundulus species, which does not appear to have active transport of

Cl- across the gills (Patrick et al., 1997; Patrick and Wood, 1999; Wood and

Marshall, 1994). The limited capacity for active Cl- uptake at the gills of

freshwater-acclimated Fundulus species places greater emphasis on reducing

paracellular Cl- loss or obtaining Cl- from dietary sources in order to maintain Cl-

balance-. . It has been proposed that development of a mechanism to minimize

paracellular Cl- loss at the gills was an important evolutionary step allowing

certain populations of SW-tolerant Fundulus to survive in FW(Laurent et al., 2006;

Scott et al., 2004b; Wood and Laurent, 2003; Wood et al., 2010; Wood and

Grosell, 2008).

Interestingly, intraspecific differences in FW tolerance appear to exist

between populations of Fundulus heteroclitus. Based on variations in

mitochondrial haplotype, nuclear proteins and microsatellites, Fundulus

heteroclitus can be divided into at least two clades: a northern and a southern

Page 25: Osmoregulation and acid-base tolerance in fish

15

clade (Bernardi et al., 1993; Gonzalez-villasenor and Powers, 1990). A

phylogenetic break in Fundulus heteroclitus appears at the Hudson River (approx.

40.5o north latitude), in which northern groups are distinct from southern groups

at mitochondrial loci (Gonzalez-villasenor and Powers, 1990; Smith et al., 1998).

These two populations exhibit variations in several physiological aspects. For

example, northern populations of Fundulus heteroclitus exhibit higher fertilization

success and survival rates in hypoosmotic exposures than do southern

populations(Able and Palmer, 1988b; Fangue et al., 2006; Scott and Schulte,

2005). It is likely that molecular or physiological differences in FW tolerance exist

between populations of Fundulus heteroclitus. A further study found that northern

populations of Fundulus heteroclitus were better adapted to FW environments

and that minimizing Cl- imbalance through regulation of paracellular pathways

appeared to be the key physiological differences accounting for their greater FW

tolerance(Scott et al., 2004b; Whitehead et al., 2011).In summary, minimizing

chloride loss through regulation of paracellular pathways is an important strategy

applied by Fundulus species in order to cope with hypoosmotic challenges, which

is the issue I will address below.

Significance of Paracellular Pathway in Osmoregulation

Ion transport across the paracellular pathways is controlled by tight junction

proteins (Tang and Goodenough, 2003)(Fig 1.3). The paracellular pathway is

regulated by a series of tight junction proteins, including occludins, claudins, and

junctional adhesion molecules, that help demarcate the environmentally-exposed

apical membrane from the blood-exposed basolateral membrane of transporting

epithelia (Chiba et al., 2008; Van Itallie and Anderson, 2004). Amongst the

Page 26: Osmoregulation and acid-base tolerance in fish

16

members of this protein complex, the claudin superfamily of proteins has

received much attention recently due to their importance in regulating the

permeability and ion selectivity of the paracellular pathway Claudins, which are

part of the tetraspanin superfamily of transmembrane proteins found in all

vertebrate epithelia, contain two extracellular loop domains and constitute the

permeability barrier to limit diffusion of solutes between cells (Koval, 2006). The

pair of hairpin domains from one cell is able to dimerize with the extracellular

domains of an adjacent cell forming a continual seal. Recent studies have

demonstrated that claudin proteins are the key components regulating the

permeability and the ion selectivity of the paracellular pathway (Koval, 2006; Van

Itallie and Anderson, 2004, 2006). It is plausible to infer that claudins play key

roles in osmoregulation in Fundulus species because Cl- concentration in plasma

in Fundulus is mainly maintained by decreasing passive loss through paracellular

pathways as described above especially when food source is limited. Some

studies have shown that some claudin proteins tend to form „tighter‟ tight junction

pores, while others are more likely to form „leakier‟ tight junction pores

(Alexandre et al., 2007; Sas et al., 2008). Claudins typically form homopolymers,

but can also form heteropolymers in certain situations (Koval, 2006; Van Itallie

and Anderson, 2004).

Until now at least 24 claudins have been found in mammals (Morin, 2009).

The claudin family seems to have expanded along chordate lineage due to

genome duplication in some fish species with 56 claudins identified in Fugu

rubripes (Loh et al., 2004). Some studies have suggested that the expansion of

Page 27: Osmoregulation and acid-base tolerance in fish

17

claudin family may due to the unique challenges of fish residing in an aquatic

environment which force fish deal with more environmental stressors(Bagherie-

Lachidan et al., 2009). In fish, preliminary studies have shown that the mRNA

and protein levels of certain claudins in transporting epithelia are responsive to

changes in environmental salinity (Bagherie-Lachidan et al., 2008, 2009).

Although fish and mammals share many claudin orthologs, roles of claudins may

Figure 1.3: A typical model showing the structure of tight junction proteins in paracellular pathways between two adjacent cells(Fromm and Schulzke, 2009)

be different between fish and mammals. For example, claudin-7 mRNA

expression does not respond to salinity challenges in Fundulus grandis (data not

shown), while one study showed that overexpression of claudin-7 decreased

paracellular Cl- conductance in LLC-PK1 cells (Alexandre et al., 2007). There is

only very limited study indicating claudin distribution among different epithelial

Page 28: Osmoregulation and acid-base tolerance in fish

18

cell types in fish (Tipsmark et al., 2008a; Tipsmark et al., 2008b; Tipsmark et al.,

2008c). Considering two major cell types, MR cells and pavement cells (PV),

expressed in fish gills, it is reasonable to infer that claudin proteins may vary

among the junctions between MR-MR, MR-PV and PV-PV cells (Phuong et al.,

2009).

General Principles of Acid-base Tolerance in Teleost Fish

In recent years, there has been significant progress in elucidating the

mechanisms of ion and acid-base regulation in FW fish. Numerous studies have

described a linkage between the exchange of acid-base equivalents and strong

ions in FW-adapted teleost fish. The „strong ion difference theory‟ was first

described by Stewart as a tool to access acid-base status independently of the

water titration method (Stewart, 1983). Briefly, both Na+ and Cl- are the major

strong cation and anion that have appreciable fluxes across the branchial

membrane. A disparity between Na+ and Cl- next fluxes usually indicates a

charge imbalance which in FW fish has been equated to the net flux of acidic

equivalents (ie., JH+net = JCl

net – JNanet) (Kirschner, 1997; Patrick et al., 1997;

Wood et al., 1984). The correlation has been reported between JH+net which is

measured by titration and the difference between the net Na+ and Cl- fluxes

(Patrick et al., 1997). A greater rate of Cl- loss over Na+ loss that is measured

following HCl injection represents a net H+ excretion.

The relationship between acid-base regulation and osmoregulation has

been demonstrated by various studies among different teleost fish species. In

this model, Na+ and Cl- uptake in FW fish was initially suggested to be through

electro-neutral Na+/H+ (NH4+) and Cl-/HCO3

- exchangers, respectively. This

Page 29: Osmoregulation and acid-base tolerance in fish

19

plausible explanation by assuming the existence of two exchangers went

unchallenged for decades. It is not until recently that some studies indicated that

this assumption of linkage between acid-base tolerance and osmoregulation is

not universal and apparently could not be applied to all fish species.(see below)

Some investigated this issue in Fundulus species in early 1990‟s and found that

the mechanisms of ion and acid-base regulation in FW-adapted Fundulus

heteroclitus departed from the standard model for FW-adapted teleosts (Patrick

et al., 1997; Patrick and Wood, 1999).One study pointed out clearly in these

studies that anion permeability of paracellular pathways would be selectively

elevated in the case of HCl-induced acidosis, which facilitates Cl- excretion

through paracellular pathways(Patrick et al., 1997). These studies uncover the

importance of paracellular pathways in Fundulus species during the recovery

from metabolic acidosis and that the regulation of paracellular ion efflux rates

play important roles in acid-base regulation, which makes measuring Na+ /Cl- flux

rates good parameters to understand the acid-base regulatory

mechanisms(Evans, 2006; Goss et al., 1992; Patrick and Wood, 1999; Wood and

Marshall, 1994).

Fundulus species are unique amongst most teleost in their response to

disturbances in the acid-base status of their extracellular fluids. These

differences include: 1) The non-existence of Cl-/HCO3- exchanger, which puts the

necessity of alternative pathways for base secretion; and 2). The relative high

flux rates of Na+ across branchial epithelia in FW fish,. Fundulus species could

modulate Na+ flux rates across paracellular pathways as a way to regulate H+

Page 30: Osmoregulation and acid-base tolerance in fish

20

excretion from the body. Briefly, Na+ and Cl- are the major strong ions that have

considerable flux rates across the branchial membrane. If a disparity between

Na+ and Cl- flux rates exists, a charge imbalance which has been equated to the

net acidic equivalents in FW fish may arise. This charge imbalance may the

compensated by H+ transport which is a part of acid-base regulation in

fish(Patrick et al., 1997; Wood et al., 1984).

Based on some previous studies in Fundulus heteroclitus, acid-base

regulatory process involves both transcellular and paracellular transport. In

addition, paracellular transport seems to play more important roles compared to

other fish species considering that a Cl-/HCO3- exchanger may not exist in

Fundulus species. These discrepancies encourage us to further examine ion and

acid-base regulation mechanisms and the nature of their linkage in FW-adapted

Fundulus heteroclitus as well as the role of the regulation of paracellular

permeability which is closely related to ion transport and acid-base regulation. As

discussed above, claudin proteins are major components of regulating ion

transport and permeability in paracellular pathways, previous studies showed

different claudins may exhibit different permeability to ions. A change in

permeability to Na+ versus Cl- may attribute to the accomplishment of an osmotic

or acid-base regulatory process. We put special emphasis on studying the

potential roles of claudins in the mechanisms of acid-base regulation in Fundulus

species.

Brief Outline of Main Points of Each Chapter

Despite these recent advances, our knowledge of mechanisms of

osmoregulation and acid-base tolerance in euryhaline fish is still far from

Page 31: Osmoregulation and acid-base tolerance in fish

21

complete. In my dissertation, I will address some of these questions utilizing

whole animal, cellular, and molecular approaches in order to reach a better

understanding of the underlying mechanisms of osmoregulation and acid-base

tolerance in euryhaline Fundulus species.

Chapter 2: This chapter aims to uncover the important roles of branchial

paracellular pathways in fish osmoregulation. Overall, the collected results

demonstrate that Fundulus grandis maintains hydromineral balances by rapidly

regulating gill permeability to certain ions. More specifically, branchial Cl-

permeability decreased significantly as a mechanism to maintain plasma Cl-

concentration. This decreased permeability is likely caused by some claudin

proteins which regulate the ion transport across paracellular pathways.

Chapter 3: This study aims to initiate investigating the mechanisms of acid

tolerance among three populations of Fundulus species. In addition, claudin

expression is measured to study if some claudins contribute to acid tolerance in

two populations of Fundulus heteroclitus. The collected data show Fundulus

majalis does not have the ability to regulate against metabolic acidosis, whereas

Fundulus heteroclitus are able to compensate quickly metabolic acidosis by

regulating the relative transports of Na+ and Cl-. As an investigation to study if

intraspecific differences of acid-base regulation exists among Fundulus

heteroclitus as that of osmoregulation does, our data show that there are no

differences in responses to acid challenge between two populations of Fundulus

heteroclitus in this study. There are no differences in responses to acid

challenge between two populations of Fundulus heteroclitus. Adaptation to acidic

Page 32: Osmoregulation and acid-base tolerance in fish

22

challenge is detected in both populations of Fundulus heteroclitus. Differential

claudin gene expression is detected and may contribute to this regulatory

process.

Page 33: Osmoregulation and acid-base tolerance in fish

23

CHAPTER 2: ALTERATIONS IN CLAUDIN EXPRESSIONS IN THE GILLS OF FUNDULUS GRANDIS FOLLOWING ABRUPT SALINITY TRANSFER

Introduction

Osmoregulation in euryhaline fish involves the coordinated regulation of

both transcellular and paracellular pathways in ion-transporting epithelia such as

the gills, intestines, and kidney (Evans et al., 2005; Marvao et al., 1994;

Shehadeh and Gordon, 1969; Tipsmark et al., 2010). The gill is particularly

important in mediating these ion movements due to the extensive surface area of

the tissue in direct contact with the external environment, and because of the

high water and blood flow rates continually passing across their extracellular

surfaces(Evans, 2008; Hwang and Lee, 2007; Wilson and Laurent, 2002). In sea

water (SW), the fish gill actively excretes ions to the external environment to

balance the passive ion gain occurring at the gills and intestines. In contrast, the

freshwater (FW) fish gill actively absorbs ions to compensate for a passive ion

loss to the more dilute environment (Chang et al., 2001; George et al., 2006b;

Patrick and Wood, 1999). Although the physiology of transcellular ion transport

has been studied for nearly a century, comparably little is known regarding the

mechanistic basis of paracellular ion movement in fish (Evans, 2008; Evans et al.,

2005).

The paracellular pathway is regulated by a series of tight junction proteins,

including occludins, claudins, and junctional adhesion molecules, that help

demarcate the environmentally-exposed apical membrane from the blood-

exposed basolateral membrane of transporting epithelia (Chiba et al., 2008; Van

Itallie and Anderson, 2004). Amongst the members of this protein complex, the

Page 34: Osmoregulation and acid-base tolerance in fish

24

claudin superfamily of proteins has received much attention recently due to their

importance in regulating the permeability and ion selectivity of the paracellular

pathway (Koval, 2006; Van Itallie and Anderson, 2006). Claudins are tetraspan

proteins, which each have a pair of extracellular domains that can form functional

dimers with the domains on neighboring cells to form charge and size selective

pores (Koval, 2006). There are at least 24 claudin orthologs in mammals (Van

Itallie and Anderson, 2006), and at least 56 claudins in Fugu fish, due to genome

duplication in fishes (Bagherie-Lachidan et al., 2009; Loh et al., 2004).Each

claudin may assume specific functions which are different from those of other

claudins(Alexandre et al., 2007; Angelow and Yu, 2007; Coyne et al.,

2003b).Expansion of the claudin gene family may be due to the response to the

unique aquatic physiological environments in fish (Bagherie-Lachidan et al.,

2009). The large numbers of claudin family contribute to completing complex

regulations of ion transport across paracellular pathways. Two claudins integrate

together as a dimer which functions as a complete unit, which makes elucidation

of the functions of claudins complicated (Koval, 2006; Morita et al., 1999).

Though there are some claudin orthologs comparable to mammals, the

expansion of the claudin family strongly suggest that claudin may behave

differently in the function of osmoregulation between mammals and fishes.

Recent studies have shown a correlation between claudin expression

changes and salinity acclimation in euryhaline fish. For example, up-regulation of

claudin-28a and claudin-30 were found associated to the FW fish gill phenotype.

to contribute to in tilapia and a similar observation has been made for claudin-27

Page 35: Osmoregulation and acid-base tolerance in fish

25

in eel (Goggisberg and Hesse, 1983; Tipsmark et al., 2008a; Tipsmark et al.,

2008b). Studies have shown that two claudin proteins form a complete functional

unit which shows specific permeability to different ions (Alexandre et al., 2007;

Alexandre et al., 2005; Angelow and Yu, 2007). This characteristic plays

important roles in regulating ion movement across paracellular pathways in

epithelia. However most knowledge of claudins obtained until now are from

studies in mammals. Little has been investigated on the distributions and

functions of claudins in fish (Bagherie-Lachidan et al., 2008, 2009; Clelland et al.,

2010; Loh et al., 2004). Considering the changing environments in fish habitats,

the study on potential functions of claudins in fish will contribute much to our

understanding on mechanisms and regulation of epithelial barriers.

Killifish do not exhibit active Cl- uptake in FW and solely rely on efflux

manipulation to maintain ion balance during hypoosmotic exposures when

dietary sources of Cl- is limited (Patrick et al., 1997; Patrick and Wood, 1999).

This efflux manipulation is achieved by adjusting gill permeability which occurs

via paracellular pathways (Patrick et al., 1997; Wood and Marshall, 1994).

Northern populations of Fundulus heteroclitus survived better following FW

challenge than southern populations during FW exposures because they

exhibited lower branchial permeability which lead to less Cl- loss through

paracellular pathways (Scott et al., 2004b). Thus, understanding the regulation of

branchial permeability of paracellular pathways seems to be the key in

uncovering the mechanisms of hypoosmotic tolerances in killifish.

Page 36: Osmoregulation and acid-base tolerance in fish

26

The aim of this study was to investigate the roles of paracellular pathways

during hypoosmotic tolerances of the euryhaline Fundulus grandis. We examined

the plasma Na+ and Cl- concentrations and ion flux rates, as well as gill

paracellular permeability in fish before and after the hypoosmotic exposures.

Fundulus heteroclitus and Fundulus grandis are two genetically closely-related

species with similar morphologies and ecological niches(Duggins et al., 1989;

Hsiao and Meier, 2005). We obtained 6 claudin-like ESTs of Fundulus

heteroclitus and designed primers that successfully worked in Fundulus grandis.

We examined mRNA expression of these claudin-like genes in different cell

phenotypes in Fundulus grandis gills using quantitative PCR. This study is the

first study of the roles of paracellular pathways in Fundulus grandis and is the

first to investigate differential expression of claudins in different cell types of gills

in fish during hypoosmotic challenges.

Materials and Methods

Experimental Animals

Adult Fundulus grandis (weight range 3.1-16.8 g) were obtained from a

local hatchery (Gulf Coast Minnows, Thibodeaux, LA) and acclimated in 330-liter

glass aquaria for at least one month prior to experimentation. Fish were housed

and handled according to an approved Institutional Animal Care and Use

Committee Protocol in facilities managed by the Department of Laboratory

Animal Medicine, Louisiana State University and accredited by the Association

for Assessment and Accreditation of Laboratory Animal Care International. Fish-

holding tanks were part of a recirculating system that received water after

filtration through separate particle and biological filters, and sterilization through

Page 37: Osmoregulation and acid-base tolerance in fish

27

an ultraviolet light. Water salinity and temperature, which were monitored daily,

were maintained at 5 ppt and 22-24oC, respectively, and nitrogenous waste

products, which were measured at least three times a week, were all kept at

negligible levels. Fish were fed twice daily at a total ration of 2% body-weight per

day with commercial killifish pellets (Cargill Aquaxcel™) and kept on a 12 h light:

12 h dark cycle.

Tissue Sampling

One hundred and fifty fish were divided randomly into 5 groups and

transferred to 100 liter glass aquaria containing 5 ppt, 2 ppt, 1 ppt, 0.5 ppt and

0.1 ppt water, respectively. These water were prepared by adding sea salt to RO

(Reverse Osmosis) water with the calibration of a salinity meter. Water samples

of 0.1 and 0.5 ppt were taken daily for determination of water chemistry (Table

2.1), water samples of 5, 2, 1 ppt were calibrated by a salinity meter daily. Fish

were sampled at 6 h, 24 h, 3 days, 7 days and 14 days post transfer (n= 6 fish

per treatment); fish were not fed for at least 12 h prior to sampling. All the fish

were net captured and anesthetized briefly in 0.5g l-1 tricaine methanesulfonate

(MS-222). Blood was collected using micro-hematocrit capillary tubes (Fisher

Scientific, USA) after severing the spine. Blood was then centrifuged at 3000 rpm

for 3 minutes, and blood plasma collected and stored at -20oC awaiting analyses

of plasma sodium and chloride concentration. Plasma sodium concentrations

were measured by flame atomic absorption spectroscopy (Varian Australia Pty

Ltd, Australia) and plasma chloride concentrations were measured using a

modified mercuric thiocyanate method (Zall et al., 1956). Gill baskets were

removed completely from each fish and washed with deionized water for 10 s.

Page 38: Osmoregulation and acid-base tolerance in fish

28

After the wash, the first left gill arch of each gill basket was processed as

described below. The remaining gill basket was cut into small pieces and

immersed into RNAlater (Invitrogen, USA) and stored at 4oC for 16 h, before

being transferred to -20oC for longer term storage.

Table 2.1 Ion concentration (mM) of the waters used in the experiments.

Salinity [Na+] [Cl-] [Ca2+] [Mg2+]

0.1 ppt 0.61±0.04 1.70±0.09 0.40±0.04 0.17±0.02

0.5 ppt 6.93±0.59 8.21±0.67 2.10±136 0.69±0.08

1 ppt 14.62 17.03 0.32 1.67

2 ppt 29.20 34.10 0.64 3.34

5 ppt 73.12 85.15 1.60 8.35

Values are represented as the mean ± the standard deviation of the mean; n=7, n is the number of water samples measured at each salinity. Ion concentrations of 1, 2, 5 ppt treatments are estimated based on the proportion of the measured ion concentrations of full strength SW (32ppt).

Sodium and Chloride Flux Experiments

Sodium and chloride flux experiments were performed using a protocol

similar to that described previously (Patrick and Wood, 1999). Thirty Fundulus

grandis weighing from 4.16 to 7.12 g were acclimated to a 38-liter 5 ppt water

tank for two weeks prior to the flux experiments and transferred to 0.5-liter of 0.1

ppt water. On the day before the experiment, six fish were transferred to six

individual chambers containing 250 ml of aerated water at 0.1 ppt. The chambers

Page 39: Osmoregulation and acid-base tolerance in fish

29

were covered to decrease fish stress, which appeared to be minimal within 10

min after transfer based on resting status of the fish after initial agitation. Fish

were maintained in the dark for 1 h until 0.5 µCi 22Na+ (PerkinElmer, USA) was

added to each chamber. Pre-experiments showed 1 h in dark was enough to

eliminate fish stress which is determined by mobile agitation of the fish in the

chamber and the deviation of flux rates from control status. Our data showed that

there were no significant differences on flux rates before and 1 hour after the

transfer (data not shown). The water was well mixed before taking water samples.

Two ml water samples were taken at 0, 0.5, 1, 2, 4, and 6 h from the time

radioisotopes were added to the water. After the experiments, fish were gently

washed, weighed and put back to a 30-liter tank with aerated 0.1 ppt water. The

same procedure was repeated at 24 h, 3 days, 7 days and 14 days after the first

flux experiments using the same fish. Fish were fed with commercial killifish food

(Cargill Aquaxcel™) twice a day, although no food was given the morning prior to

the start of flux experiments. The radioactivity of 22Na+ was counted on a Gamma

Radiation Counter (Gamma Counter 5500, Beckman Instruments Inc, USA) and

sodium concentration in water samples was measured by flame atomic

absorption spectroscopy (Varian Australia Pty Ltd, Australia).

The chloride flux experiments were performed by calculating the rates of

disappearance of chloride radioactive isotope from the water and the

accumulation rates of isotopes in fish body, which was modified from a protocol

previously described (Patrick and Wood, 1999). One ml of freshwater containing

0.25 µCi 36Cl- was added to each chamber containing 200 ml of 0.1 ppt water.

Page 40: Osmoregulation and acid-base tolerance in fish

30

Each 1 ml water sample taken during the experiment was added to 5 ml Ultima

Gold scintillant (PerkinElmer Life and Analytical Sciences, USA) and the solution

was kept in the dark overnight for full interactions. The samples were then

counted on a scintillation counter (TRI-CARB 2900TR Liquid Scintillation

Analyzer, PerkinElmer Life and Analytical Sciences, USA). All samples had

similar quench, so therefore, no quench correction factor was required.

The unidirectional influx rates for Na+ and Cl- were calculated using the

change in water isotope radioactivity between successive time points as

calculated by:

Jin=

Eq. 1

where cpm1 and cpm2 are the radioisotope activities at the start and end of each

time period, volume is the exact volume of water in the flux chamber, weight is

the fish weight in grams, and SA is the mean specific activity of the water in

cpm/µmol.

Net flux rates were calculated by the total ion concentration differences

among water samples by the following equation:

Jnet =

Eq. 2

where [ion1] and [ion2] are the ion concentrations at the start and end of flux

period respectively. All other variables are similar to those described for Equation

1.

Efflux rates were calculated by the differences between influx and net flux

rates.

Page 41: Osmoregulation and acid-base tolerance in fish

31

Je=Jnet - Jin Eq. 3

Only net flux rate were data presented for the 3 day, 7 day, and 14 day time

points due to undetectable Cl- influx rates at 3, 7, 14 day time point after fish

exposed to 0.1 ppt water.

SEM and TEM

Fish gills were fixed in 2% glutaraldehyde-1% formaldehyde in 0.2 M

cacodylate buffer for 4 h, then rinsed 5X in 0.1 M cacodylate buffer containing

0.02 M glycine over 12 h period. Samples were post-fixed in 2% osmium

tetroxide for 1 h, then rinsed in water, en bloc stained in 5% uranyl acetate in the

dark for 1 h, rinsed in water 2X, and dehydrated in an ethanol series which is a

step-by-step 20-minute dehydration from 30% ethanol to 50%, 70%, 95% and

100% ethanol sequentially. The samples were critical-point dried with liquid CO2

in a Denton CPD, mounted on aluminum SEM stubs, coated with gold: palladium

at a ratio of 60: 40 in an Edwards S150 sputter coater, and imaged with a JSM-

6610 high vacuum mode SEM.

Emersion of live killifish in lanthanum (La) nitrate emersion was used as a

method to differentiate between leaky and tight paracellular tight junction

complexes in the fish gill. Briefly, fish were immersed for 15 minutes into a 1.5%

solution of LaNO3 dissolved in 5, 2, 1, 0.5, 0.1 ppt water, respectively.

Lanthanum went into paracellular pathways if it was leaky while lanthanum could

not if it was tight. Lanthanum ion turned into black in TEM images after a series

of treatment addressed below. Before fish transfer, the transfer medium was

alkalinized to pH 7.8 to avoid precipitation of La hydroxide. Following exposure,

fish were sacrificed, and the gills were dissected from the fish and fixed in 2%

Page 42: Osmoregulation and acid-base tolerance in fish

32

glutaraldehyde-1% formaldehyde in 0.2 M sodium cacodylate buffer. The

samples were rinsed in fixative for 10 seconds at least 3 times to remove

external remnants of lanthanum nitrate. The samples were then cut into small

pieces and treated with osmium for 1 h. After dehydration with ethanol, the gill

filaments were treated for transmission electron microscopy (TEM) imaging taken

at apical sides of branchial epithelia at magnifications of 26000X and 260000X by

the Advanced Microscopy Center in the Department of Biological Sciences at

Louisiana State University.

Gill Permeability Experiments: Tritium-labelled polyethylene glycol

(3[H]PEG-4000; MW=4000, 18.5 MBq/g, American Radiolabeled Chemicals, Inc)

is a high molecular weight polymer of ethylene oxide and is a blend of polyers

with different degrees of polymerization. It is very soluable to water and widely

used as an indicator of epithelial paracellular permeability. Gill permeability was

calculated in previous studies using different methods (Kumai et al., 2011; Scott

et al., 2004b; Wood et al., 1998). Seventy-two Fundulus grandis were acclimated

to 5 ppt water for at least two weeks and then randomly assigned into three

groups and transferred to 5, 0.5 and 0.1 ppt water for 0, 1, 3, or 7 days. At each

of these time points, six fish per salinity were transferred to individual, darkened,

flux chambers containing 1.25 liter of water at the same salinity they had been

exposed to post transfer (i.e., 5 ppt, 0.5 ppt, and 0.1 ppt). After a 1 h acclimation

period, 25 µCi 3[H]PEG-4000 was added to each container and two water

samples (1 ml each) were taken at time 0 and 6 h after addition of 3[H]PEG-

Page 43: Osmoregulation and acid-base tolerance in fish

33

4000.Fish were weighted and sacrificed at the end of each 6-hour exposure

period.

Fish intestine was gently removed carefully to ensure no purging of

intestinal contents. Each intestine and its associated gut contents were digested

in 1M HNO3 overnight at 60oC and the digested content was centrifuged at 4000

rpm for 5 min with an aliquot of the supernant diluted to a v:v ratio of 1:2 with

liquid scintillant (Ultima Gold scintillant liquid, Perkin Elmer Life and Analytical

Sciences, USA). The 1 mL water samples were diluted with 4 mL deionized

water, and diluted to a v:v ratio of 1:2 with liquid scintillant. The radioactivity of

these samples was counted using a scintillation counter (TRI-CARB 2900TR

Liquid Scintillation Analyzer, PerkinElmer Life and Analytical Sciences, USA).The

parameter of gill paracellular permeability is the clearance of external radioactive

isotopes by calculating the appearance of radioactivity within fish body minus

uptake of that in intestine versus time. This method excludes drinking effect

which represents uptake of radioactivity in intestine and thus is more accurate

than traditional methods(Scott et al., 2004b).

Gillp=

Eq. 4

where [isotopes in fish body]a represents amount of isotopes in fish body at

the end of experiment, [isotopes in intestine]a represents the amount of isotopes

in intestine at the end of the experiment. Time represents duration of the

experiment. Weight represents fish weight at the end of the experiment.

Page 44: Osmoregulation and acid-base tolerance in fish

34

Total RNA Isolation

Total RNA from gill tissue was isolated using TRIzol reagent (Invitrogen,

Carlsbad, California, USA) according to the manufacturer‟s instructions and total

RNA concentrations measured using a NanoDrop spectrophotometer (NanoDrop

Technologies, Wilmington, Delaware, USA). Single-stranded cDNA synthesis

(using 2µg total RNA per 20 µL reaction) was primed at poly(A) tail by reverse

transcription using reverse transcription kit (High Capacity cDNA Reverse

Transcription Kit, Applied Biosystems, Carlsbad, California, USA).

Claudin mRNA Levels in Isolated Gill Cells

Forty-eight Fundulus grandis were acclimated to 5 ppt water for at least two

weeks, then divided into two groups and transferred to 5 or 0.1 ppt water. Fish

(n=6 per salinity) were sampled after 6 h, 24 h, 3 days, and 7 days post transfer.

Gill epithelial cells were isolated from each fish using a protocol previously

described (Galvez et al., 2002). Briefly, epithelial cells from each fish were

separated into different cell types by using a three-step Percoll gradient

consisting of 1.03, 1.06, 1.09 g.ml-1 Percoll in a similar protocol. After the

isolation, 300 nM fluorescent Mito-tracker Red was applied to cells obtained from

different layers to check percentages of MR cells in each fraction to ensure

proper cell type separation. Mitochondrion-rich cells (MRCs) displayed extensive,

strong fluorescent staining within cell boundaries while non-mitochondrion-rich

cells only showed sparse and weak fluorescent staining inside the cells. After

the separation of the cells from each layer, I counted MRC number in each field

of view under fluorescent microscopes. MRCs consist of 87-95% (91.6%±0.02) of

total cells in 1.06-1.09 layer while that is only less than 5% (4%±0.017) in 1.03-

Page 45: Osmoregulation and acid-base tolerance in fish

35

1.06 layer. Total RNA was isolated fromepithelial cells collected from 1.03-1.06

and 1.06-1.09 g.ml-1 layers using Trizol reagents. RNA was converted to single-

stranded cDNA (using 2µg total RNA per 20 µL reaction) by reverse transcription

(High Capacity cDNA Reverse Transcription Kit, Applied Biosystems, Carlsbad,

California, USA).

Quantitative PCR

Fundulus heteroclitus claudin-like EST sequences (DR047164, DN951669,

CV819612, DR441634, DR441341, DR441481) were obtained using the National

Center for Biotechnology Information (NCBI). These ESTs were submitted to a

BLAST search against the non-redundant nucleotide database and assigned the

nomenclature, claudin-3, claudin-5, claudin-7, claudin-23, claudin-26, and

claudin-28, respectively, based on their highest sequence similarity to Takifugu

rubripes claudin nucleotide sequences. These EST sequences were used to

develop primers for qPCR using Primer 3 software (Primer3) ensuring that a GC

base content between 40-60%, a theoretical annealing temperature of 60 C, and

a primer length of approximately 21 nucleotides were maintained. All primers

amplified single products as demonstrated by agarose gel electrophoresis and

denaturation analysis following individual real-time PCR (qPCR). Primer

sequences are listed in Table 2.2.

The qPCR analysis was performed using a SYBR Green detection system

(SYBR Green core Reagent, Applied Biosystems, USA) according to the

manufacturer‟s instructions. Reactions were carried out with 5 µL cDNA, 10 nM

forward and reverse primers, and 2 µL SYBR Green reagent in a total volume of

20 µL. A three-step cycling protocol was used with 40 cycles of 50 oC for 2 min,

Page 46: Osmoregulation and acid-base tolerance in fish

36

95oC for 10 min, and 95 oC for 15 s. Critical threshold (Ct) values were calculated

using the adaptive baseline function in the ABI Prism 7000 SDS software (ABI

Prism 7000, Applied BioSystems). The amplification efficiencies for each primer

set was calculated by serially-diluting cDNA derived from the gills of control fish,

and using these dilutions to obtain a linear regression relationship of Ct. For all

primer sets, amplification efficiency varied between 92.1% and 105.5%, with 100%

efficiency representing a theoretical doubling in the amount of cDNA after each

cycle.

Relative changes in mRNA levels were calculated by the 2 -∆∆Ct method

using the formula, ΔCt target (treatment-control) - ΔCt ref (treatment-control)

(Galvez et al., 2007), where target refers to the claudin genes, and ref refers to

the reference gene, 18s ribosomal RNA (18s rRNA). Controls represent the 5.0

ppt-acclimated fish at each time point, and treatments represent either the 0.1

ppt or 0.5 ppt fish under hypoosmotic challenges. The Ct values for 18s rRNA

varied only moderately among the different treatments with a standard error of

0.057 within each plate of samples plate. The mean ΔΔCt value of each 5.0 ppt

control was expressed as 1, and relative change in mRNA levels of hypoosmotic

treatments were expressed relative to this value. A terminal dissociation step (95

oC for 15 s, 60 oC for 20 s, and 95 oC for 15 s) was added to verify the presence

of only one amplicon in the PCR reaction.

Statistical Analysis

All data collected were expressed as mean ± SE (n) where n equals the

number of fish in an experimental group. Plasma chemistry data were analyzed

by one-way ANOVA with Student-Newman-Keuls test. After the test of

Page 47: Osmoregulation and acid-base tolerance in fish

37

homogeneity of variance, gill permeability data were analyzed by one-way

ANOVA. Equal variance was checked on the data of claudin expression levels in

different cell types before a two-way ANOVA statistical method was applied to

test the significance of claudin expressional levels among 5, 0.5, and 0.1 ppt

samples. A significance level of P<0.05 was chosen. All statistics were done by

SPSS version 17.

Table 2.2 Sequences of forward and reverse primers used to measure Fundulus grandis mRNA claudin levels using quantitative PCR.

Name Direction Sequence 5’3’

Cldn3-qPCR Forward AAGCTTCTCTTGACCCAGCA

Cln3-qPCR Reverse CCCACCACTGCGAGACTAAT

Cldn5-qPCR Forward TGGGACTTGCTATGTGCGTA

Cldn5-qPCR Reverse ACGGAGGAGATGATGGTGAG

Cldn7-qPCR Forward CATCATCCTGATCGGAGCTT

Cldn7-qPCR Reverse TGGCACCTCCAATTATAGCC

Cldn23-qPCR Forward CGAACAAACAACAACCATGC

Cldn23-qPCR Reverse GAAGATCCCTTGCTCGATGA

Cldn26-qPCR Forward GGACTGCAGATCGTTGGTTT

Cldn26-qPCR Reverse TGGCCACCATCACTGATCTA

Cldn28-qPCR Forward TCTGGAGTTGCCTCAAGACC

Cldn28-qPCR Reverse ATGATTGTGTTGGCAGTCCA

18s-qPCR Forward TTCCGATAACGAACGAGAC

18s-qPCR Reverse GACATCTAAGGGCATCACAG

Page 48: Osmoregulation and acid-base tolerance in fish

38

Results

Plasma Chemistry

One hundred and fifty-six Fundulus grandis were acclimated to 5ppt water

for at least 2 weeks before they were divided randomly and transferred to 5, 2, 1,

0.5, 0.1 ppt water for hypoosmotic challenges respectively. Plasma from each

sacrificed fish was collected, then [Na+] and [Cl-] of these samples were

measured as the direct parameters of fish responses to hypoosmotic challenges.

Plasma [Na+] significantly decreased 6 h after the transfer to 0.1 ppt water,

reached its lowest point at 24 h and tended to recover thereafter, but could not

recover completely even 14 days after the transfer. There was only a significant

difference at 6 h after the transfer in 0.5 ppt group. No significant differences

were found at 24 h, 3, 7 and 14 days (Fig 2.1A).

Plasma [Cl-] data showed significant decreases at all time-points after the

transfer at 0.1 ppt with the lowest point at 6 h compared to control water (5 ppt).

No significant differences found in 0.5 ppt group (Fig 2.1B). Plasma [Na+] and [Cl-]

did not show disturbances within 1, 2 and 5 ppt groups.

Unidirectional Sodium and Chloride Fluxes

Changes in sodium and chloride flux rates give direct exhibition of

regulations in sodium and chloride ion movement across fish epithelia during

hypoosmotic challenges.

During the first 6 h after Fundulus grandis were transferred from 5 ppt to

0.1 ppt, the average sodium influx rate was 260 µmol-1.kg-1.h-1 and the efflux rate

was approximately at -770 µmol-1.kg-1.h-1. There was a significant net efflux rate

Page 49: Osmoregulation and acid-base tolerance in fish

39

during the first 6 h. The influx rates remained relatively constant 1 day and 3

days after the transfer while the efflux rates decreased significantly at 1 day and

*

**

**

*

-2 0 2 4 6 8 10 12 14

Pla

sm

a [

Na

+]

(mM

)

140

150

160

170

180

190

200

0.1 ppt

0.5 ppt

1.0 ppt

2.0 ppt

5.0 ppt

*

**

*

A

Time Post-transfer (Days)

222

*

**

**

*

-2 0 2 4 6 8 10 12 14

Pla

sm

a [

Cl- ]

(mM

)

80

100

120

140

160

0.1 ppt

0.5 ppt

1.0 ppt

2.0 ppt

5.0 ppt*

*

*

*

B

Time Post-transfer (Days)

222

** *

Figure 2.1 Plasma sodium (A) (N=6) and chloride (B) (N=6) levels in Fundulus grandis before and after transfer from 5 ppt water to 2 ppt, 1 ppt, 0.5 ppt and 0.1

ppt water respectively. Data are expressed as means ± S.E.M. * indicates

significance from pre-transfer 5 ppt control (P<0.05).

3 days and no significant net flux rates were found. Both influx and efflux rates

increased significantly 7 days after the transfer but the net flux rate was not

significantly different from zero. There was no difference in efflux rates between

first 6 h and 14 days. After 14 days post transfer, the influx rate increased to

Page 50: Osmoregulation and acid-base tolerance in fish

40

approximately 730 µmol.kg-1.l-1 with the efflux rate at 760 µmol.kg-1.l-1. No net flux

was detected 14 days after the transfer (Fig 2.2A).

The average chloride influx rate in the first 6 h after the transfer to 0.1 ppt

0-6 h 1 d 3 d 7 d 14 d

Na

+ F

lux (

µm

ol kg-1

h-1

)

-1000

-750

-500

-250

0

250

500

750

1000Influx

Efflux

Net flux

(solid bar)

A

a a a

b

b

a'

a''

b''

b''b''

b''

b'b'

c'

a'

0-6 h 1 d 3 d 7 d 14 d

Cl- F

lux (

µm

ol kg-1

h-1

)

-1250

-1000

-750

-500

-250

0

250

500

750

B

a

a

a'

a''

b''

b''b'' b''

b'

Influx

Efflux

Net flux

(solid bar)

Figure 2.2: Total (net) flux and unidirectional fluxes of Na+ (A) and Cl- (B) in Fundulus grandis after transfer from 5 ppt water to 0.1 ppt water (N=6) versus time. Positive values represent influx (Jin), negative values represent efflux (Jout). a, b were used to indicate the significant differences in influx rates; a‟, b‟, c‟ were used to indicate the significant differences in efflux rates; a‟‟, b‟‟ were used to indicate the significant differences in net flux rates.

Page 51: Osmoregulation and acid-base tolerance in fish

41

water was 250 µmol-1.kg-1.h-1 with the average efflux rate at -790 µmol-1.kg-1.h-1. A

significant net efflux rate of -540 µmol-1.kg-1.h-1 was found during the first 6 h. By

1 day, no difference of influx rate was found while there was a significant

decrease in efflux rate. From 3 days post-transfer, influx and efflux was so

minimal that it was difficult to detect them. As discussed earlier, Fundulus

species do not have active chloride uptake systems in FW. Minimizing chloride

loss is the strategy to maintain plasma chloride levels. This strategy seemed to

happen 3 days after FW exposure. Only net flux rate was measured during this

period but no significant difference was detected (Fig 2.2B).

Scanning Electron Microscopy (SEM) Study

The branchial epithelial morphology exhibits distinct differences between a

FW fish and a SW one. SEM study was applied in this experiment to give direct

evidence showing this transition of gill morphology in fish under hypoosmotic

challenges. Our data found that gills in Fundulus grandis at 5 ppt water showed

typical „SW‟ type morphology. Pavement cells (PVCs) occupied more than 90%

of surface area and mitochondrion-rich cells (MRCs) represented a small crypt at

apical sides (Fig 2.3). After 6 h post-transfer to 0.1 and 0.5 ppt, there was a

transient morphological change in MRCs which showed that cell membrane

surfaces were more exposed to the external environment with the appearance of

micro villi. These cells looked closer to typical „FW-type‟ MRCs. The apical

surfaces were square, triangular or rounded. The population of this type of cells

was about 20% that of PVCs. PVCs showed more foldings on the cell surface

compared to 5ppt group. After 1 day, FW-type MRCs in 0.1 ppt water fish

Page 52: Osmoregulation and acid-base tolerance in fish

42

continued showing this morphological change but the population of these types

of cells decreased. However morphology of gills in 0.5 ppt water fish tended to go

back to „SW-type‟ morphology 1 day after the transfer. No apparent differences

including surface areas, shape of edges in PVC epithelia were found on the

morphology of PVCs throughout the experiments at all salinity and time points

(Fig 2.3).

Figure 2.3 Gill SEM images of Fundulus grandis transferred from 5 ppt to 0.1ppt and 0.5ppt group at 6 h, 24 h, 3 days and 7 days. PV represents pavement cell. MR represents mitochondrion-rich cell which are exhibited as apical black crypts . Each bar represents 5 µm.

Gill Permeability Study

Gill permeability is the characteristic showing the difficulty of an ion passing

through epithelia. This study used these data as an indicator explaining changes

in sodium and chloride flux rates. The rate of disappearance of tritium-labeled

polyethylene glycol (3[H]PEG-4000; MW=4000) across the gills did not show

significant change in the Fundulus grandis 5 ppt control group basing on a

statistical analysis among control groups. This rate did not show significant

difference at 6 h in the 0.5 ppt group compared to the 5 ppt control group,

Page 53: Osmoregulation and acid-base tolerance in fish

43

however it did decrease significantly at 1, 3 and 7 days post-transfer. The rate of

3[H]PEG-4000 disappearance showed significant decrease at all time-points in

0.1 ppt group compared to 5 ppt control (Fig 2.4) These data showed that fish

exhibited prompt responses of regulating paracellular permeability to

hypoosmotic challenges and that the lower the salinity was, the faster the

responses were.

0-6 h 1 d 3 d 7 d

Perm

eabili

ty(m

l kg

-1h

-1)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.45ppt

0.5ppt

0.1 ppt

**

** * * *

Figure 2.4 Total gill permeability rate (ml kg-1 h-1, previously described in equation 4 in Materials and Methods part) of PEG-4000 in Fundulus grandis after transfer from 5 ppt water to 5 ppt, 0.5 ppt and 0.1 ppt water respectively. Gill paracellular permeability represents the rate of external solution entering fish body by calculating the appearance of radioactivity within the fish body minus uptake of the radioactivity in intestine versus time. Data are expressed as means ± S.E.M. * indicates significant difference from 5 ppt control (P<0.05).

Transmission Electron Microscopy Study

The use of lanthanum nitrate provides useful information on differentiating

the “open” and “close” of paracellular pathways. Lanthanum went into

paracellular pathways if it was leaky while lanthanum could not if it was tight.

Lanthanum ion turned into black in TEM images after a series of treatment. Our

data showed lanthanum could penetrate into paracellular pathways between

MRCs and adjacent cells in 5, 2, 1 ppt group. However an apparent penetration

was not observed in paracellular pathways in 0.5 and 0.1 ppt groups (Fig 2.5).

Page 54: Osmoregulation and acid-base tolerance in fish

44

Figure 2.5 Junctions between MRCs and adjacent cells in Fundulus grandis 3 days after the transfer from 5 ppt water to 5 (A, a), 2 (B,b), 1 (C,c), 0.5 (D,d), 0.1 (E,e) ppt water. A (a), B (b), C (c): lanthanum penetrated paracellular pathways between mitochondrion-rich cells and adjacent cells. D (d), E (e): lanthanum could not penetrate paracellular pathways between mitochondrion-rich cells and adjacent cells. Each black box in A, B, C, D, E contains a paracellular pathway with a magnification of 26000X. The picture (a,b,c,d,e) on the right showed the paracellular pways at a magnification of 260000X Scale bars represent 1 µm.

Gene Identification and Claudin Expression Profiles

We first examined expressional patterns of these 6 claudins in gills of

Fundulus grandis following hypoosmotic exposures. Transfer of 5 ppt-acclimated

Page 55: Osmoregulation and acid-base tolerance in fish

45

Fundulus grandis to lower salinity induced different claudin expression patterns in

fish gills.

6 h 1 d 3 d 7 d

0

10

20

30

40*

A

Rela

tive m

RN

A L

evels

**

Claudin-3

6 h 1 d 3 d 7 d

0

1

2

3

4

55ppt

0.5ppt

0.1 ppt

*

D

Rela

tive m

RN

A L

evels

*

Claudin-28

6 h 1 d 3 d 7 d

0

1

2

3

4

5*

B

Rela

tive m

RN

A L

evels

Claudin-5

**

*

6 h 1 d 3 d 7 d

0

1

2

3

4

5

E

Rela

tive m

RN

A L

evels

Claudin-7

6 h 1 d 3 d 7 d

0

1

2

3

4

5

*

C

Rela

tive m

RN

A L

evels

Claudin-23

**

6 h 1 d 3 d 7 d

0

1

2

3

4

5

F

Rela

tive m

RN

A L

evels

Claudin-26

Figure 2.6 Changes of claudin expressional levels in response to hypoosmotic challenges versus time. The effects of hypoosmotic exposure (0.1 ppt and 0.5 ppt exposure) on the abundance of Cldn3 (A), Cldn5 (B), Cldn23 (C), Cldn28 (D), Cldn7 (E) and Cldn26 (F) in gills of Fundulus grandis for 6 h, 1 day, 3 days and 7 days. Relative claudin mRNA expression was determined relative to controls with no hyposomotic challenge. All expressional changes were normalized to expression of EF1a(we used both, we first used 18s a little, then we used EF1a) A T-test was used to assess the expressional changes of claudins in gills of Fundulus grandis between 5ppt control and 0.5 ppt, 0.1ppt experimental groups (P<0.05).

Page 56: Osmoregulation and acid-base tolerance in fish

46

Claudin-3 expression showed no difference at 6 and 24 h after 0.1 ppt

transfer. A robust 24-fold expression increase was detected by 3 days and then

decreased by 7 days. Claudin-3 expression in 0.5 ppt group was more

progressive with an 8-times increase found by 7 days. However claudin-3

expression levels showed discrepancies between MRCs and PVCs. Claudin-3

expression showed a sharp increase 24 h in MRCs after the transfer and

remained at high expression levels thereafter while only a slow increase was

detected in PVCs 3 days after the transfer (Fig 2.6A, 2.7A).

Claudin-5 increased 3.6 folds 24 h after the exposure to 0.1 ppt water and

remained significantly different from control group at 3 and 7 day time point. No

significant differences detected in 0.5 ppt group until 7 days after the transfer.

Increases of claudin-5 expression occurred exclusively in MRC 24 h after the

exposures, whereas no expressional change was detected in PVCs throughout

the experiment. These findings confirmed that claudin-5 was mainly expressed in

MRCs instead of PVCs. The data also indicated that claudin-5 may be a

functional part contributing to osmoregulatory mechanisms in fish during

hypoosmotic exposures. (Fig 2.6B, 2.7B).

Claudin-23 showed an increase in both 0.1 and 0.5 ppt group 3 days after

the exposure, however this increase persisted only in 0.1 ppt group 7 days after

the exposure. Expression of claudin-23 increased 6 h after the exposure and this

up-regulation persisted across the experiments in MRCs. Only a transient

increase 24 h after the exposure was detected in PVCs. These finding indicated

hypoosmotic changes of the environments. (Fig 2.6C, 2.7C).

Page 57: Osmoregulation and acid-base tolerance in fish

47

6h 1d 3d 7d

Re

lative

mR

NA

Le

ve

ls

0

5

10

15

20

25

5 ppt

0.1 ppt

6h 1d 3d 7d

A

**

*

*

*

PVC MRC

Claudin 3

6h 1d 3d 7d

Re

lative

mR

NA

Le

ve

ls

0

5

10

15

20

25

6h 1d 3d 7d

B

*

*

*PVC MRC

Claudin 5

6h 1d 3d 7d

Re

lative

mR

NA

Le

ve

ls

0

3

6

9

12

6h 1d 3d 7d

C

*

**

PVC MRC

*

*

Claudin 23

6h 1d 3d 7d

0

1

2

3

4

5

6h 1d 3d 7d

D

* *

PVC MRC

Claudin 28

6h 1d 3d 7d

0

1

2

3

4

5

6h 1d 3d 7d

E

PVC MRC

Claudin 7

6h 1d 3d 7d

0

1

2

3

4

5

6h 1d 3d 7d

F

PVC MRC

Claudin 26

5 ppt

0.1 ppt

Figure 2.7 The effects of hypoosmotic exposure (0.1 ppt exposure) on the abundance of Cldn3 (A), Cldn5 (B), Cldn23 (C), Cldn28 (D), Cldn7 (E) and Cldn26 (F) in two different epithelial cells: PVCs and MRCs for 6 h, 1 day, 3 days and 7 days. Relative claudin mRNA expression was determined relative to controls with no hypoosmotic challenge. A T-test was used to assess the expressional changes of claudins in the same type of cells between 5ppt control and 0.1ppt experimental group (P<0.05).

that claudin-23 was mainly expressed in MRCs and was responsive to

Claudin-28 expression showed an abrupt decrease 3 days after transfer to 0.1

ppt which is correlated to the expressional decrease in PVCs. These data

suggested that claudin-28 was mainly expressed in PVCs. No expressional

Page 58: Osmoregulation and acid-base tolerance in fish

48

change was detected in MRCs (Fig 2.6D, 2.7D). Claudin-7 and claudin-26 did not

show expressional changes in both experiments, which suggested both were

unresponsive to hypoosmotic challenges and may assume a function of structure

in paracellular pathways. (Fig 2.6E, 2.7E, 2.6F, 2.7F).

Discussion

The gulf killifish, Fundulus grandis, and its sister taxa the Atlantic killifish,

Fundulus heteroclitus, reside in coastal marshes that are subject to frequent and

episodic oscillations in salinity, dissolved oxygen, and temperature (Marshall,

2003; Wood and LeMoigne, 2001).Their ability to tolerate these dynamic habitats

has made them a popular model to study the physiological basis of

environmental stress tolerance (Burnett et al., 2007a). Although tending to inhabit

brackish to marine environments (Kaneko and Katoh, 2004; Wood and Grosell,

2008, 2009), euryhaline killifish are exposed to reduced environmental salinity

while feeding, or during natural or anthropogenic freshwater inputs such as

rainstorm flooding events (Marshall, 2003). Killifish appear to cope with short

bouts of fresh water by making rapid physiological adjustments (Marshall, 2003;

Wood and Grosell, 2008, 2009); it is only with prolonged exposure to fresh water

that fish are then required to make more substantive compensatory adjustments

to their ion-transporting epithelia (Wood and Grosell, 2008).

Few studies have attempted to delineate the salinity at which the gill

epithelium transitions from a SW type to a FW phenotype. Interestingly, most

Fundulus species are ostensibly marine teleost with the ability to retain their SW

physiology at salinities approaching FW (Feldmeth and Waggoner, 1972; Griffith,

1974a; Haag and Warren, 2000). Copeland and Philpott (Copeland, 1950;

Page 59: Osmoregulation and acid-base tolerance in fish

49

Philpott and Copeland, 1963) suggested killifish make this physiological transition

at a salinity between 2.0 and 0.5 ppt; however no concrete data were provided in

these studies to support this claim. Recently, we have shown that Fundulus

heteroclitus acclimated to full-strength SW can tolerate abrupt transfer to 0.4 ppt

with no significant effects to plasma osmolality over 14 days post-transfer

(Whitehead et al., 2010). In the present study, killifish were acclimated to 5.0 ppt

since we had previously shown that the gill surface morphology retains its SW

phenotype at this salinity, and since we wanted to test this SW to FW phenotypic

transition over a relatively small salinity range. Interestingly, we found that

plasma Na+ and Cl- concentrations were only significantly decreased at salinities

of 0.5 ppt and below, and that only exposure to 0.1 ppt elicited a sustained

reduction in plasma Cl- concentration from pre-transfer controls (Fig 2.1). This is

in accordance with our scanning electron microscopy data (Fig 2.3) showing that

the gills of Fundulus grandis transition to a FW phenotype occurs at a salinity of

0.1 ppt. Although the gills of fish exposed to 0.5 ppt expressed features

consistent with a FW phenotype after only 6 h post exposure, the apical crypts

typical of a SW-acclimated fish gill reemerged by 1 d. The surface morphology of

mitochondrion-rich gill cells in 0.1 ppt-exposed killifish is consistent with the

surface characteristics of the cuboidal cell first described by Laurent et al

(Laurent et al., 2006); cells analogous to the mitochondrion-rich PV cells that are

presumed to have high active Na+ transport capacity but only limited ability to

actively transport Cl-(Laurent et al., 2006). These data also support the presence

of fully-developed MR cell types beneath the gill surface, which can be quickly

Page 60: Osmoregulation and acid-base tolerance in fish

50

expressed on the surface upon osmotic challenge (Katoh et al., 2001), and a gill

that transitions from a SW to a FW phenotype at approximately 0.5 ppt.

Maintenance of osmotic balance in euryhaline fish during hypoosmotic

exposure requires both a reduction of paracellular loss and a longer term

stimulation of solute absorption in existing or newly-formed cells (Lin et al., 2004;

Marshall et al., 2005). These physiological processes can be monitored indirectly

using unidirectional solute movement in whole animals, which acts as a proxy of

gill function. Unidirection influx and efflux rates of Na+ in Fundulus heteroclitus

ranged from 15000-18000 µmol·kg-1·h-1 in 100% SW and 3000 µmol·kg-1·h-1 in

10% SW, but decreased to below 1000 µmol·kg-1·h-1 during FW exposure (Wood

and Marshall, 1994). In the present study, Na+ influx rate was approximately 250

µmol·kg-1·h-1 and Na+ efflux rate was approximately 750 µmol·kg-1·h-1 in

Fundulus grandis after 6 hours exposure to 0.1 ppt water (i.e., FW) (Fig 2.2). This

net Na+ imbalance recovered to normal (i.e., net Na+ flux approaching 0) after 1

day exposure as a result of reduction in Na+ efflux over this acute time frame.

Na+ influx remained relatively constant until day 3 post transfer to 0.1 ppt, but

then gradually increased thereafter, indicative of an increase in active Na+ uptake

via a transcellular pathway. This stimulation of active Na+ influx paralleled a

secondary, concomitant increase in Na+ efflux such that net Na+ balance

remained negligible. One study found that Fundulus heteroclitus could not

maintain sodium balance in their bodies when the concentration of sodium in

external environments was below 0.98mM which was called balance point(Wood

and Marshall, 1994). The sodium concentration of 0.1 ppt water in this study was

Page 61: Osmoregulation and acid-base tolerance in fish

51

less than 0.7mM. Two reasons may explain this contradiction: 1. Fundulus

grandis has a lower balance point compared to Fundulus heteroclitus. 2.

Fundulus grandis was fed during the experiments in this study so dietary

compensation of sodium may help animal maintain sodium levels in fish plasma.

Cl- influx and efflux rates were both below 1000 µmol·kg-1·h-1 in Fundulus

grandis 6 hours after exposure to 0.1 ppt water, but decreased to a negligible

rate by day 3 post exposure (Fig 2.2).One study reported that Cl- infux and efflux

rates were above 15000 µmol·kg-1·h-1 in SW-adapted Fundulus heteroclitus

(Wood and Marshall, 1994), suggesting a rapid and dramatic decrease in both

unidirectional rates upon transfer to FW. Previous studies have demonstrated

that Fundulus heteroclitus have no appreciable active Cl- uptake in FW, and

suggest that regulation of paracellular Cl- loss is a key strategy of hypoosmotic

tolerance in FW-tolerant Fundulus populations (Scott et al., 2004b). Some

studies have reported that euryhaline teleost transferred from SW to FW were

able to regulate plasma osmolality within days, but often did so at an osmolality

lower than that pre-transfer (Kalujnaia, 2007). This is consistent with the plasma

Cl- concentrations of the 0.1 ppt fish observed here, which stabilized by day 3

post-transfer, but at concentrations significantly below those of the 5.0 ppt control

fish. These finding indicated that animal may lower chloride concentration in

plasma as another strategy to cope with hypoosmotic challenges besides the

strategy to minimize passive chloride loss.

Wood and Grosell (2008; 2009) reported that transepithelial potential,

which is an important determinant of the electrochemical driving force of gill ion

Page 62: Osmoregulation and acid-base tolerance in fish

52

transport, was relatively refractory to the short-term transfer of SW-acclimated

Fundulus heteroclitus to FW, and that only with longer exposure to FW did the

gills of Fundulus heteroclitus functionally switch into that of a FW animal (Wood

and Grosell, 2008, 2009). The gills of brackish water-acclimated Fundulus

exhibited high non-selective cation permeability largely via the paracellular shunt

pathways (Wood and Grosell, 2008, 2009). Brackish water-acclimated Fundulus

heteroclitus exhibited a positive diffusional transepithelial potential (TEP), which

facilitates Na+ extrusion via paracellular shunt pathways (Wood and Grosell,

2009). The TEP of the Fundulus in natural environments is interpreted as a

diffusional potential dictated by the relative permeability of the branchial epithelia

to the two dominant ions (Na+ and Cl-) in extracellular fluid and the external

environments (Wood and Grosell, 2008). Upon acute exposures to FW, the TEP

will become negative and limits Na+ loss immediately. However this negative

TEP in plasma side does not hamper the uptake of Cl- due to the fact there is no

active Cl- uptake in the gills of Fundulus species (Patrick and Wood, 1999; Wood

and Marshall, 1994). Regardless, keeping Cl- loss to negligible amounts but

maintaining extensive paracellular permeability to Na+ is a unique feature of the

gills of Fundulus species in FW.

Numerous studies have shown that claudins, a key constituent of the tight

junction complex, are the principal proteins forming barrier properties regulating

permeability of paracellular pathways(Koval, 2006; Van Itallie and Anderson,

2004, 2006). Although not well studied, there is growing evidence that claudins

are likely important regulators of paracellular ion movement in the fish gill, and

Page 63: Osmoregulation and acid-base tolerance in fish

53

are under dynamic regulation during osmotic challenges (Bagherie-Lachidan et

al., 2009; Clelland et al., 2010; Tipsmark et al., 2008a). A transcriptional change

in the abundance of specific claudins or claudin isoform switching could

contribute to the transformation of the fish gill from a „leaky‟ epithelium to a „tight‟

one in euryhaline fish during hypoosmotic challenges.

Although the ability of gills to remodel with varying environmental salinity is well

reported (Katoh et al., 2001; Laurent and Dunel, 1980; Wilson and Laurent,

2002), few investigations have described the cellular mechanisms regulating their

transient and long-term responses to osmotic stress. The current study

demonstrated significant increases in the mRNA levels of claudin 3, 5, and 23 in

the gills of Fundulus grandis during hypoosmotic exposures, while that of claudin-

28 decreased following FW exposure (Fig 2.6). Claudin-3 protein levels in the

gills of tilapia have been shown to increase following transfer from SW to FW

(Tipsmark et al., 2008a), which is in accordance with the increase in mRNA

levels in this study. It is possible that claudin-3 is involved in forming „tighter‟

paracellular pathways which favor FW tolerance. We also found that claudin-3

mRNA levels increased in both PVCs and MRCs (Fig 2.7), suggesting this

isoform may be associated with PVC-PVC and PVC-MRC tight junctional

complexes in FW-acclimated killifish, and that it might associated with the

increased “tightening” of the gill epithelium in FW. Acclimation of Fundulus

heteroclitus to FW is known to elicit the proliferation of a cuboidal-type MR cell,

this study also found that the mitosis rate of these „cuboidal cells‟ reached the

peak at 12 h after the exposure (Laurent et al., 2006). This increased mitosis rate

Page 64: Osmoregulation and acid-base tolerance in fish

54

may correlate with increased expression of some claudins which are mainly

expressed in these cuboidal cells. Our data found the increased expression of

claudin-3 occurred 24 h after the transfer. It is plausible to infer that this sharp

up-regulation may be probably correlated with these newly expressed „cuboidal

cells‟. These „cuboidal cells‟ , which show different morphology from typical FW-

type MRCs which exhibit active Cl- uptake as previously described in teleost

(Evans et al., 2005), are possibly responsible for the lack of Cl- uptake in

Fundulus species (Laurent et al., 2006).

Claudin-5 mRNA expression levels increased significantly in Fundulus

grandis gills following hypoosmotic exposure, however, unlike the situation for

claudin-3 mRNA expression, transcript abundance was highest in the MR cell

fraction (Fig 2.6, Fig 2.7). A recent study indicated that claudin-5 contributed to

forming a „tight‟ brain barrier in zebrafish (Abdelilah-Seyfried, 2010), which

contrasts the reported stimulatory effect on paracellular permeability of this

claudin in mammals (Wen et al., 2004). Interestingly, one study showed that a

significant portion of SW-type MRCs were replaced by FW-type MRCs following

3 days transfer from SW to FW (Katoh and Kaneko, 2003). In the current study,

claudin-5 mRNA levels in the 0.1 ppt gills were significantly increased above that

of the 5 ppt controls at 24 h post transfer, and reached levels of 10-fold above

controls by 3 d and 7 d post transfer. It is possible that this increased claudin-5

expression may contribute to forming a “tighter” paracellular pathway and be

related to long-term adaptation of hypoosmotic challenges.

Page 65: Osmoregulation and acid-base tolerance in fish

55

Claudin-23 mRNA levels were significantly increased after hypoosmotic

exposure and this increase was also predominately localized to the MR cell

fraction (Fig 2.6, Fig 2.7), which indicated its role in FW tolerance. These are the

first set of data reported so far outlining a potential role of this claudin isoform in

fish osmoregulation. Interestingly, this increase in claudin-23 expression occurs

as early as 6 h after the transfer. Taken that morphological transformation occurs

at 6 h in the SEM study (Fig 2.3) into consideration, these data suggest that

claudin-23 expression increase may relate to short-term transformation in MRCs

in Fundulus grandis which was described in another study earlier (Katoh and

Kaneko, 2003). There is also an increase of claudin-23 in PV cell fraction 3 days

after the transfer, which may possibly be due to contamination during cell

separation.

In contrast, claudin-28 mRNA levels decreased significantly in Fundulus

grandis following FW exposure (Fig 2.6, Fig 2.7). However, this decrease in

transcript abundance was observed in the PV cell fraction, suggesting a possible

role of PVCs, and not only MR cells, in hypoosmotic tolerance. Interestingly,

another study did not show any effect of salinity transfer on claudin-28 mRNA

levels in Atlantic salmon (Tipsmark et al., 2008b), suggesting species-specific

differences of this claudin in response to osmotic response. The absence of

claudin 28 in mammals suggests that this isoform may confer unique attributes to

the paracellular pathway in fish. Our data did not detect claudin-7, -26 expression

changes in Fundulus grandis with hypoosmotic challenge. One study found that

overexpression of claudin-7 decreased paracellular Cl- conductance and

Page 66: Osmoregulation and acid-base tolerance in fish

56

increased paracellular Na+ conductance in LLC-PK1cells (Alexandre et al.,

2005).These controversial findings suggest that further investigation is needed in

order to reach a better understanding.

Despite these expressional changes of claudins described above, it is

important to point out that the nomenclature of the claudins in this study may not

be very accurate. Due to extensive gene duplication in fish compared to

mammals, existence of subunits of claudins is prevalent (Loh et al., 2004). This

study however does give valuable information on the relation of claudins and FW

tolerance in Fundulus species, which is the first one investigating on this question.

We used radiolabeled-PEG 4000 as a marker of paracellular permeability

to provide insights in the possible function of claudins during acclimation to

hypoosmotic conditions. Previous studies have administered radiolabeled PEG-

4000 into fish by intraperitoneal injection and then monitored the appearance of

the marker in water. This method is based on the assumption that the

transcellular movement of PEG-4000 was negligible (Kumai et al., 2011; Scott et

al., 2004b). However, despite of this assumption, PEG-4000 is able to permeate

kidney glomeruli and enter into water via urination (Scott et al., 2004b), which

makes estimation of extrarenal routes of PEG-4000 permeation problematic. As

such, we decided to add PEG-4000 to the water and measure its appearance

into the animal, after removing any contribution associated with drinking. Our

data found that amount of PEG taken into gastrointestinal tract accounted for 2-

21% of total amount of PEG into the body in Fundulus grandis (data not shown),

which indicated effect of drinking is substantial in FW Fundulus grandis. . Thus

Page 67: Osmoregulation and acid-base tolerance in fish

57

by applying PEG into external medium and excluding the effect of drinking, the

method we used in this study gave more accurate measurement on evaluating

paracellular permeability. Our data showed an acute reduction of paracellular

permeability occurs promptly after the hypo-osmotic exposure. In addition, we

recorded abrupt decrease in both Na+ and Cl- efflux rates. It is likely that the

sharp decrease of Na+ and Cl- efflux rates were caused by the decrease of

paracellular permeability, which was significant in minimizing ion loss in the early

stage of post-transfer period and maintaining plasma ion balance. It is surprising

to discover that branchial paracellular permeability did not show differences

between 0.5 ppt and 0.1 ppt group 24 h after the transfer in this study (Fig 2.4).

In order to reach a better understanding of these data, we applied lanthanum

staining in TEM study depicting the paracellular pathways. We found there was

no lanthanum penetration into the paracellular pathways in 0.5 and 0.1 ppt group

while significant penetration into paracellular pathways in 5, 2 and 1 ppt group

(Fig 2.5). This finding gave strong evidence that Fundulus grandis shut down

paracellular pathways to minimize passive ion loss during hypoosmotic

exposures. However this robust decrease in paracellular permeability is unlikely

to be the result of structural transformation or protein expression changes which

usually takes hours or days to happen. Some studies indicated that acute

regulation of claudins by phosphorylation may play a role in regulation of the

paracellular permeability (Angelow and Yu, 2007). This is not surprising since

claudins are the major components of tight junctions regulating paracellular

permeability (Koval, 2006). Phosphorylation has been linked to both increases

Page 68: Osmoregulation and acid-base tolerance in fish

58

and decreases in tight junction assembly and function. There are some studies

indicating that a group of kinases like protein kinase A and myosin light chain

kinase are involved in the regulation of claudins or other membrane proteins

(Koval, 2006), however this regulatory process is still unclear and further

investigation is needed.

In summary, three strategies were adopted by Fundulus species during

acclimation to hypoosmotic exposure basing on this and previous studies: 1.

Decreasing Na+ loss by switching the TEP from positive to negative (Wood and

Grosell, 2009); 2 Minimizing Na+ and Cl- loss by decreasing permeability of

paracellular pathways; 3 Adjusting plasma [Na+] and [Cl-] at a lower level to

lessen the stress of ion maintenance. In addition, this study is the first to give

experimental evidence on the transition point from „SW morphology‟ to „FW

morphology‟ in fish osmoregulation. The information of claudin expressions in

this experiment will contribute to a better understanding on their roles in fish

osmoregulation. In addition, this study gave valuable information on the roles of

claudins during hypoosmotic exposures in Fundulus grandis. Further work needs

be done on exploring the function of the whole family of claudins in terms of fish

osmoregulation from protein levels to reach a better understanding of the

underlying mechanisms.

Page 69: Osmoregulation and acid-base tolerance in fish

59

CHAPTER 3: ACID-BASE TOLERANCES AMONG FUNDULUS SPECIES

Introduction

Under freshwater conditions, teleost fish are hyperosmotic to their

environment, and must actively absorb osmolytes such as Na+ and Cl- to

compensate for the passive loss of ions through paracellular pathways (Evans

and Claiborne, 2006). One interesting feature is that the transport of strong ions

in plasma(Patrick et al., 1997), such as Na+ and Cl- in freshwater teleost relies on

the counter movement (i.e., excretion) of acid and base equivalents into the

surrounding environment as a requirement for the maintenance of electro-

neutrality(Marshall and Grosell, 2005). Compared to animals living on land, fish

have a limited capacity for respiratory compensation of non-respiratory, acid-

base disturbances, largely due to physical constraints of living in an aquatic

environment (Evans and Claiborne, 2006). Although changes in renal excretion

and intercellular buffering may contribute to their regulation, ion and acid-base

homeostasis primarily occurs at the gills (Evans and Claiborne, 2006; Evans et

al., 2005).

The gill epithelium is a dynamic ion-transporting epithelium that manipulates

the rates of sodium and chloride transport to achieve both ion and acid-base

homeostasis(Evans, 2008; Hwang and Lee, 2007; Wilson and Laurent, 2002).

Numerous studies have described a linkage between acid-base tolerance and

transepithelial ion exchange in the gills of FW-acclimated fish (Evans et al., 1999;

Hyde and Perry, 1987; Marshall and Grosell, 2005; Patrick and Wood, 1999;

Salama et al., 1999; Wilson et al., 1995; Wood and Marshall, 1994). In most

species, Cl- is removed from the water in exchange for a basic equivalent, HCO3-,

Page 70: Osmoregulation and acid-base tolerance in fish

60

whereas Na+ is removed from the water in exchange for an acidic equivalent, H+

(Goss et al., 1992; Heisler, 1984). Although considerable controversy exists

regarding the mechanism of Na+ and H+ exchange in the gills, most fish species

have a direct coupling of Cl- and HCO3- exchange in the gill, through a putative

anion exchanger on the apical membrane of the tissue. Recent studies, however,

suggest that Fundulus species are amongst a few teleost known to have little to

no active transport of Cl- at the gills (Patrick and Wood, 1999). As a result,

several studies have implicated the restriction of paracelluar Cl- loss as a

requirement for hypoosmotic tolerance; a mechanism that likely delineates

species and/or population tolerance of fresh water in fish from the genus

Fundulus. Furthermore, lack of this transport system likely imparts differences in

the regulation against acid-base disturbances.

One study found that metabolic acidosis in Fundulus heteroclitus initiated by

the intraperitoneal injection of HCl had no effect on Na+ or Cl- influx rates, but

rather stimulated an increase of Cl- loss and attenuation of unidirectional Na+

efflux (Patrick et al., 1997). This study drew on the possible role of regulation of

paracellular loss of ions as a mechanism of acid-base regulation (Evans et al.,

2005; Goss and Wood, 1990a; Goss and Wood, 1990b; Patrick et al., 1997;

Patrick and Wood, 1999; Wood and Marshall, 1994). These studies highlight the

importance of the paracellular pathway and the associated tight junction proteins

in acid-base regulation.

The claudins are a group of proteins which span bilayers 4 times. Some

studies show claudins are the key proteins in regulating paracellular permeability

Page 71: Osmoregulation and acid-base tolerance in fish

61

and solute transport (Koval, 2006). Claudins of a given cell interact with other

claudins in the adjacent cells and two claudin proteins form a functional dimer

which is charge- selective or size-selective (Angelow and Yu, 2007; Coyne et al.,

2003a; Coyne et al., 2003b; Van Itallie and Anderson, 2006; Wen et al., 2004).

Until now at least 24 claudins have been found in mammals (Angelow and Yu,

2007; Koval, 2006). The claudin family seems to have expanded along chordate

lineage due to genome duplication in fishes, with a total of 56 claudins identified

in Fugu fish (Loh et al., 2004). Some conjectured that the expansion of claudin

family may due to the response to the unique aquatic physiological environments

in fish (Bagherie-Lachidan et al., 2009). Though there are some claudin orthologs

comparable to mammals, the expansion of the claudin family strongly suggest

that claudin may behave differently in the function of osmoregulation between

mammals and fishes.

Fundulus heteroclitus populations exhibit both physiological and adaptative

responses to cope with the changing environments they inhabit (Burnett et al.,

2007b). This study aims to uncover underlying mechanism of acid-base tolerance

by comparing responses to acid challenge of two Fundulus species: Fundulus

heteroclitus and Fundulus majalis. A special effort will be put on two populations

of Fundulus heteroclitus to study if intraspecific variation of acid tolerance exists.

In addition, some claudin expressions will be investigated to study if there is any

potential claudin involved in acid-base tolerances between two populations of

Fundulus heteroclitus since previous studies suggested that claudins play key

roles in regulating paracellular ion transport(Koval, 2006).

Page 72: Osmoregulation and acid-base tolerance in fish

62

Materials and Methods

Experimental Animals

Fundulus heteroclitus (weight range of 3.8-9.6 g) were collected from

Piscataway Park, Maryland (termed Fundulus heteroclitus MDPP) and

Chincoteague National Wildlife Refuge, Chincoteague Island, Virginia (termed

Fundulus heteroclitus VAcoast) during June 2011, and transferred to Louisiana

State University, where they were held for at least 1 month. Fundulus heteroclitus

MDPP were held at 0.1 ppt, which was approximately the salinity that they were

collected from in the field. Fundulus heteroclitus VAcoast were collected from

water at 32 ppt, but maintained in the laboratory at 18 ppt, until approximately

one month prior to experimentation, when they were gradually transferred to 0.1

ppt over the course of a week, then acclimated at this salinity for the duration of

experimentation. Fundulus majalis VAcoast ,a typical marine killifish, were

collected from Chincoteague National Wildlife Refuge, Chincoteague Island,

Virginia during June 2011, and transferred to Louisiana State University, where

they were held in seawater originally and gradually brought down to 0.1 ppt

freshwater over a course of two weeks.

All populations were acclimated for at least one month at 0.1 ppt in 250-liter

glass aquaria. Fish were housed according to an approved Institutional Animal

Care and Use Committee Protocol in facilities managed by the Department of

Laboratory Animal Medicine, Louisiana State University and accredited by the

Association for Assessment and Accreditation of Laboratory Animal Care

International. Fish-holding tanks were part of a recirculating system that received

water after filtration through separate particle and biological filters, and

Page 73: Osmoregulation and acid-base tolerance in fish

63

sterilization through an ultraviolet light. Prior to experimentation, water salinity

and temperature, which were monitored daily, were maintained at 0.1 ppt and 22-

24oC, respectively, and nitrogenous waste products, which were measured at

least three times a week, were all kept at negligible levels. Fish were fed twice

daily at a total ration of 2% body-weight per day with commercial killifish pellets

(Cargill Aquaxcel™) and kept on a 12 h light: 12 h dark cycle.

Tissue Sampling and Flux Experiment

Thirty fish per population (Fundulus heteroclitus MDPP and Fundulus

heteroclitus VAcoast; weighing from 3.8 to 9.6 g) which were already acclimated

to 0.1 ppt water for at least one month were injected by intraperitoneal injection

with 1000 nEq.g-1 HCl (using a stock of 140 mM HCl) and immediately

transferred back to 0.1 ppt. Twenty-four of these fish were transferred to glass

aquaria containing 0.1 ppt, whereas the remaining 6 fish were used immediately

in radioisotopic flux experiments as described below. At 12h post transfer to 0.1

ppt, an additional fish (n=6) were transferred from the glass aquaria to individual

flux chambers as described below. The remaining fish (n=18) were injected from

days 1 to 7 post transfer with 1000 nEq.g-1 HCl, and utilized at 1d, 3d, and 7 days

for radioisotopic flux experiments. Six fish of Fundulus majalis VAcoast were

injected by intraperitoneal injection with 1000 nEq.g-1 HCl (using a stock of 140

mM HCl) and immediately transferred to 0.1 ppt. Six Fundulus majalis VAcoast

were injected with 1000 nEq.g-1 HCl (using a stock of 140 mM HCl) and and

immediately transferred back to 0.1 ppt water. Since this species could not

survive after the HCl injection, no further flux experiments were done during 12h,

24h, 3d, 7d time points.

Page 74: Osmoregulation and acid-base tolerance in fish

64

Sodium flux experiments were performed in opaque, plexiglass containers

containing 250 mL 0.1 ppt water, according to a protocol previously described

(Patrick and Wood, 1999). Briefly, each fish was transferred to an individual flux

chamber, and allowed to acclimate to the conditions for exactly 1 hour. . Pre-

experiments showed 1 h in dark was enough to eliminate fish stress which is

determined by mobile agitation of the fish in the chamber and the deviation of flux

rates from control status. Our data showed that there were no significant

differences on flux rates before and 1 hour after the transfer. . After 1 h, 0.5 uCi

22Na+ (as NaCl; PerkinElmer, USA) was added to each chamber and mixed well

by gentle aeration. Fish were held under static conditions for the entire 6 h flux

period. Water samples (2 mL per time point) were taken at 0, 1, 2, 3, 4, 5, 6 h

from the start of the flux period. After flux experiments (described below), all the

fish were net captured and anesthetized briefly in 0.5 g.l-1 tricaine

methanesulfonate (MS-222). Gill baskets were removed from each fish and

washed with deionized water for 10 seconds. After the wash, two gill arches of

each fish was cut into small pieces and immersed into 1 ml RNAlater (Invitrogen,

USA) solution and stored at 4oC for 16 h, then transferred to -20oC awaiting RNA

isolation as described below. Similar fluxes were repeated on fish at 12 h, 24 h, 3

days and 7 days. Fish were fed with commercial killifish food twice a day. No

food was given to fish the morning prior to the flux experiments. 22Na+

radioactivity was counted on a Gamma Radiation Counter (Gamma Counter

5500, Beckman Instruments Inc, USA). Sodium concentration of water samples

Page 75: Osmoregulation and acid-base tolerance in fish

65

was measured by flame atomic absorption spectroscopy (AA240FS, Varian

Australia Pty Ltd, Australia).

The unidirectional influx rates for Na+ were calculated using the change in

water 22Na+ radioactivity between successive time points as calculated by:

Jin=

where cpm1 and cpm2 are the radioisotope activities at the start and end of each

time period (i.e., 0-1h, 1-2h, 2-3h, 3-4h, 4-5h, and 5-6h), volume is the exact

volume of water in the flux chamber, weight is the fish weight in grams, and SA is

the mean specific activity of the water in cpm/µmol.

Net flux rates, which represent the difference between unidirectional influx

and efflux rates, were calculated using the change in total Na+ concentration in

water between successive time points according to the following equation:

Jnet=

where [ion1] and [ion2] are the total Na+ concentrations at the start and end of

each time period (i.e., 0-1h, 1-2h, 2-3h, 3-4h, 4-5h, and 5-6h).

Unidirectional efflux rate was calculated for each fish by the difference between

unidirectional Na+ influx rate and net Na+ flux rate.

Total RNA Isolation

Gill tissue total RNA was isolated by TRI reagent (Invitrogen, Carlsbad,

California, USA) with RNA quality confirmed by gel electrophoresis (1%

agarose). RNA concentration and purity were determined by spectrophotometry

using a Nanodrop spectrophotometer before and after DNase

Page 76: Osmoregulation and acid-base tolerance in fish

66

treatment(NanoDrop Technologies, Wilmington, Delaware, USA). Absorbance

values (Abs 260/280) in all samples were 1.8-2.1. Two microgram DNase-treated

total RNA was used in a cDNA systhesis by reverse transcription using random

primers with a reverse transcription kit (High Capacity cDNA Reverse

Transcription Kit, Applied Biosystems, Carlsbad, California, USA).

Quantitative PCR

Four claudin-like EST sequences of Fundulus heteroclitus (DR047164,

DN951669, DR441634, DR441481) were acquired from the National Center for

Biotechnology Information (NCBI). A BLAST search against the non-redundant

nucleotide database was processed based on their highest sequence similarity to

Takifugu rubripes claudin nucleotide sequences and these ESTs were assigned

the nomenclature, claudin-3, claudin-5, claudin-23, claudin-38, respectively.

Primer 3 software was used to develop primers based on these ESTs for qPCR

analysis later. All designed primers were ensured a GC base content between

40-60%, a theoretical annealing temperature of 60 C, and a length of

approximately 21 nucleotides maintained. A single product was obtained by

agarose gel electrophoresis after each primer amplification. These primer

sequences are listed in chapter 2 (Table 2.2).

Quantitative real-time PCR (qPCR) analysis was performed using a SYBR

Green detection system (SYBR Green core Reagent, Applied Biosystems)

according to the manufacturer‟s instructions. Reactions were carried out with 5

µL cDNA, 10 nM forward and reverse primer, and 2 µL SYBR Green reagent in a

total volume of 20 µL. A three-step cycling protocol was used with 40 cycles of 50

oC for 2 min, 95oC for 10 min, and 95 oC for 15 s (ABI Prism 7000). Critical

Page 77: Osmoregulation and acid-base tolerance in fish

67

threshold (Ct) values were calculated using the adaptive baseline function in the

ABI Prism 7000 SDS software. The amplification efficiencies for each primer set

were calculated by serially-diluting cDNA derived from the gills of control fish,

and using these dilutions to obtain a linear regression relationship of Ct versus

the natural logarithm of relative cDNA concentration. For all primer sets,

amplification efficiency varied between 92.1% and 105.5%, with 100% efficiency

representing a theoretical doubling in the amount of cDNA after each cycle.

Relative changes in mRNA levels were calculated using the 2 -∆∆Ct method

(Galvez et al., 2007) using the formula, ΔCt target (treatment-control) - ΔCt ref

(treatment-control), where target refers to the claudin genes, ref refers to the

reference gene, (18 S ribosomal RNA (18S rRNA), controls represent the 0.1

ppt-acclimated fish at each time point, and treatments represent either the 0.1

ppt HCl-injected fish. The Ct values for 18s rRNA varied only moderately among

the different treatments with a standard error of 0.057. The mean ΔΔCt value of

each 0.1 ppt control was expressed as 1, and relative change in mRNA levels of

hypoosmotic treatments expressed relative to this value. A terminal dissociation

step (95 oC for 15 s, 60 oC for 20 s, and 95 oC for 15 s) was added to verify the

presence of only one amplicon in the PCR reaction.

Statistical Analysis

All data collected are expressed as mean ± SE (n) where n equals the

number of fish in an experimental group. Statistics on flux data and gill claudin

expression was done using one-way ANOVA with Student-Newman-Keuls test. A

significance level of P<0.05 was chosen. All statistics were done by SPSS

version 17.

Page 78: Osmoregulation and acid-base tolerance in fish

68

Results

Unidirectional Sodium Fluxes

The unidirectional Na+ flux rates for both Fundulus heteroclitus populations

and Fundulus majalis were measured in order to detect potential responding

differences among fish during acidic challenges. These data showed that the

unidirectional Na+ and Cl- flux rates of fish were significantly affected by

metabolic acidosis induced by intraperitoneal injection of HCl, and the magnitude

and time-course of this response was dependent on the amount of time fish were

kept under acidotic conditions (Fig 3.1, 3.2, 3.3).

Acidosis resulted in a significant positive net Na+ flux in Fundulus

heteroclitus MDPP between 3h to 6h following acidosis (Fig. 3.1A). Similar

effects were seen in the Fundulus heteroclitus VAcoast population, except that a

significant increase in net Na+ flux was not elicited until 4 hours following acidosis

(Fig 3.2A). In both cases, the initial increase in net Na+ flux was associated with

an increase in unidirectional Na+ influx, but with no significant effect on Na+ efflux

rate (Fig 3.1A; 3.2A). These data showed both populations of Fundulus

heteroclitus exhibited similar responses to acidic challenges.

The effect of Na+ and Cl- flux rates in response to one shot of HCl infusion

ceased within 24 hours. In order to study the long-term effect of acidic challenges,

fish received continuous infusion of HCl at 12h, 24h, 3d, 7d, respectively. During

subsequent acid infusions (i.e., 1-7d of acidosis), both net Na+ flux and

unidirectional Na+ influx rates were also increased for both Fundulus heteroclitus

populations. The major difference seen following repeated acid infusions was

that the acidosis-induced increase in unidirectional Na+ influx was initiated

Page 79: Osmoregulation and acid-base tolerance in fish

69

sooner after acidosis (Fig. 3.1B), but tended to recover back to control levels by

the end of the 6h recovery period. For example, the unidirectional Na+ influx rate

in Fundulus heteroclitus MDPP that received a second HCl injection at 24h

transfer to 0.1 ppt was elevated by 3h post acidification, although it recovered to

Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na

flu

x r

ate

s (

um

ol/kg

/h)

-1500

-1000

-500

0

500

1000

1500

Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na

flu

x r

ate

s (

um

ol/kg

/h)

-1500

-1000

-500

0

500

1000

1500 *

*"

*"

*"

*B Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na

flu

x r

ate

s (

um

ol/kg

/h)

-1500

-1000

-500

0

500

1000

1500*

*"

*"

*"

*"

C

Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na

flu

x r

ate

s (

um

ol/kg

/h)

-1500

-1000

-500

0

500

1000

1500 * *

*" *"

D

Fig 3.1: Na+ influx, efflux and net flux rates in Fundulus heteroclitus MDPP at 0 h (A), 24 h (B), 3 day (C) and 7 day (D) point during a 7-day acid challenge period. Positive bars represent influx rate. Negative bars represent efflux rate. Solid bars represent net flux rate. * represents significant difference in influx rates compared to control (0 h); *'

represents significant difference in efflux rates compared to control (0 h); *" represents significant differences in net flux rates compared to control (0 h) (p<0.05).

control levels by 6h (Fig 3.1B). The lack of effect of metabolic acidosis on

unidirectional Na+ efflux meant that changes in net Na+ flux were largely

influenced by changes in influx (Fig 3.1B). In contrast, the unidirectional Na+

Page 80: Osmoregulation and acid-base tolerance in fish

70

influx rate in the Fundulus heteroclitus VAcoast populations was only increased

at 5 h post acid infusions, although the net Na+ flux was significantly increased by

3 h, as seen in Fundulus heteroclitus MDPP. Both populations did not show

significant differences in Na+ efflux rates during this period (Fig 3.2B).

Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na flu

x r

ate

s (

um

ol/kg/h

)

-1500

-1000

-500

0

500

1000

1500

*

*"

*"

*"

**

A

Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na flu

x r

ate

s (

um

ol/kg/h

)

-1500

-1000

-500

0

500

1000

1500*

*"

*"*"

B Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na flu

x r

ate

s (

um

ol/kg/h

)

-1500

-1000

-500

0

500

1000

1500

*

*"

** *

*"

*"*"

*

C

Time after Exposure

0 h 1 h 2 h 3 h 4 h 5 h 6 h

Na flu

x r

ate

s (

um

ol/kg/h

)

-1500

-1000

-500

0

500

1000

1500

*

*"

*

*"

*"

*

D

Fig 3.2: Na+ influx, efflux and net flux rates in Fundulus heteroclitus VAcoast at 0 h (A), 24 h (B), 3 day (C) and 7 day (D) point during a 7-day acid challenge period. Positive bars represent influx rate. Negative bars represent efflux rate. Solid bars represent net flux rate. * represents significant difference in influx rates compared to control (0 h); *'

represents significant difference in efflux rates compared to control (0 h); *" represents significant differences in net flux rates compared to control (0 h) (p<0.05).

Daily intraperitoneal HCl injections for 3d caused unidirectional Na+ influx rates to

increase sooner with acidification in both populations of Fundulus heteroclitus.

Net Na+ influx was significantly increased in Fundulus heteroclitus MDPP after

Page 81: Osmoregulation and acid-base tolerance in fish

71

only 2 h, but remained elevated despite there only being an increase in

unidirectional Na+ influx at 3h post acidosis. No significant differences in Na+

efflux rates were detected (Fig 3.1C). There were some discrepancies compared

to Fundulus heteroclitus VAcoast which showed significant differences in Na+

influx rates as early as 2 h. Fundulus heteroclitus VAcoast showed significant net

Na+ gain at 3 h, 4 h, 5 h, 6 h point (Fig 3.2C).

After 7 d of daily acid infusions, both populations were responding quickly

by stimulating Na+ influx and net flux by 2h, and recovering to control levels by 3h

or 4h post acidification for the Fundulus heteroclitus MDPP and Fundulus

heteroclitus VAcoast, respectively (Fig 3.1D, 3.2D). Once again, acidification did

not affect unidirectional Na+ efflux rates in either population at any time point

following administration of a metabolic acidosis treatment (Fig 3.1D, 3.2D).

In comparison with the effects seen in Fundulus heteroclitus, metabolic

acidosis elicited severe impairment of Na+ homeostasis and led to mortality in

Fundulus majalis. Unidirectional Na+ influx was not stimulated by acidification by

intraperitoneal injection with 1000 nEq.g-1 HCl (Fig 3.3). Both Na+ efflux and net

flux rates were highly negative after only 1h; an effect that persisted until their

death. These data exhibited significant differences of responses to acidic

challenges between two species of Fundulus: Fundulus heteroclitus and

Fundulus majalis. An increase of Na+ influx rates was stimulated in both

populations of Fundulus heteroclitus while this increase was not elicited in

Fundulus majalis. In addition, both populations of Fundulus heteroclitus survived

after HCl infusion while Fundulus majais VAcoast failed to survive.

Page 82: Osmoregulation and acid-base tolerance in fish

72

Influence of Metabolic Acidosis on Claudin mRNA Levels

Four claudin ESTs were selectively studied because previous investigation

(see chapter 2) showed these four claudins expressional levels changed in

response to hypoosmotic challenges. Claudin 3-like protein mRNA levels in the

gills of Fundulus heteroclitus MDPP were significantly increased at 24 h post HCl

injection, and this significant difference persisted through 7d acidification. A

similar trend was observed in the gills of Fundulus heteroclitus VAcoast following

acidication (Fig 3.4A, Fig 3.5A).

Time after Exposure

Control 1h 2h 3h 4h

Na

flu

x r

ate

s (

um

ol/kg

/h)

-3000

-2500

-2000

-1500

-1000

-500

0

500

*"

*'

*'*'

*'

*"

*"

Fig 3.3: Na+ influx, efflux and net flux rates in Fundulus majalis after HCl injection. Positive bars represent influx rate. Negative bars represent efflux rate. Solid bars represent net flux rate. * represents significant difference in influx rates compared to control (0 h); *'

represents significant difference in efflux rates compared to control (0 h); *" represents significant differences in net flux rates compared to control (0 h) (p<0.05).

Claudin 23-like protein mRNA expressions in Fundulus heteroclitus MDPP

increased significantly after 12 h, 24 h, 3 d, and 7 d acidification. In contrast,

Page 83: Osmoregulation and acid-base tolerance in fish

73

Fundulus heteroclitus VAcoast only shows these significant increases at 24 h, 3

d and 7 d time point (Fig 3.4C, Fig 3.5C).

No significant differences are detected on claudin 5-like and claudin 28-like

protein mRNA levels throughout the experiments compared to control group (Fig

3.4B, Fig 3.4D, Fig 3.5B, Fig 3.5D).

A

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0.0

0.5

1.0

1.5

2.0

**

*

B

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0.0

0.5

1.0

1.5

2.0

C

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0

1

2

3

4

5

6

*

**

*

D

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Fig 3.4: Claudin 3 (A), -5 (B), -23 (C), -28 (D) mRNA expression level changes in Fundulus heteroclitus MDPP gills during a 7-day acid challenge period. * represents significant difference in mRNA expression levels compared to control (p<0.05)

Discussion

The linkage between osmoregulation and acid-base tolerance in FW-

adapted fish has been established for decades. One study suggested that Na+

Page 84: Osmoregulation and acid-base tolerance in fish

74

was exchanged for endogenous H+ or NH4+ and Cl- was exchanged for

endogenous HCO3- (Krogh, 1938). This process was determined in the gills of

freshwater fish via independent electroneutral exchanges (Evans et al., 2005;

Marshall and Grosell, 2005). Independent manipulation of these Na+/acid and Cl-/

A

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0.0

0.5

1.0

1.5

2.0

* *

*

B

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0.0

0.5

1.0

1.5

2.0

C

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0

1

2

3

4

5

6

*

* *

D

Time

Control 0-6 h 12 h 24 h 3 d 7 d

Rela

tive m

RN

A E

xp

ressio

n L

evels

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Fig 3.5: Claudin 3 (A), -5 (B), -23 (C), -28 (D) mRNA expression level changes in Fundulus heteroclitus VAcoast gills during a 7-day acid challenge period. * represents significant difference in mRNA expression levels compared to control (p<0.05)

base exchangers were thus applied in various studies and a significant advance

in understanding of acid-base regulation was achieved in the past several

decades (Cameron, 1978; Cameron, 1984; Claiborne et al., 1997; Goss et al.,

1998; Goss et al., 1992; Heisler, 1984; Maetz, 1976; Marshall and Grosell, 2005).

Page 85: Osmoregulation and acid-base tolerance in fish

75

This model of Na+/acid and Cl-/base exchangers does not apply for all fish

species. The linkage between ion regulation and acid-base transport in Fundulus

heteroclitus appears to be very different from that in other fish species (Patrick

and Wood, 1999). The main difference between Fundulus species and other fish

species involves an active Cl- uptake system in Fundulus species which requires

further study, that is, Fundulus species does not show an active Cl- uptake while

most other teleost do. (Patrick et al., 1997; Wood and Marshall, 1994). Lack of

active Cl- uptake in Fundulus allows a unique model species which departs from

the traditional model. Intraspecific variation exists between two populations of

Fundulus heteroclitus: northern and southern populations (Scott et al., 2004b).

Based on variations in mitochondrial haplotype, nuclear proteins and

microsatellites, Fundulus heteroclitus can be divided into at least two populations:

a northern and a southern population (Adams et al., 2006; Bernardi et al., 1993;

Duvernell et al., 2008; Smith et al., 1998; Whitehead, 2009; Whitehead, 2010;

Whitehead and Crawford, 2006; Whitehead et al., 2010; Whitehead A, 2010).

Northern populations more readily tolerate hypoosmotic challenges than

southern populations due to their better ability to minimize passive Cl- loss

through branchial paracellular pathways (Scott et al., 2004b). Based on previous

studies, Fundulus heteroclitus MDPP belongs to northern population while

Fundulus heteroclitus VAcoast belongs to southern population (Adams et al.,

2006; Smith et al., 1998; Whitehead, 2009; Whitehead, 2010; Whitehead and

Crawford, 2006; Whitehead et al., 2010; Whitehead A, 2010). Furthermore, net

Na+ and Cl- flux differential is a very practical and useful means for estimating

Page 86: Osmoregulation and acid-base tolerance in fish

76

acid-base balance in vivo in that net amount of acid secreted equals to net

amount of Na+ uptake (Patrick and Wood, 1999). The above information

construct the background of this study to investigate if intraspecific variation

exists between two populations of Fundulus heteroclitus during acidic challenge

via characterizing Na+ flux rates across fish gills.

Generally both populations of Fundulus heteroclitus showed active Na+

uptake after HCl injection throughout the experimental period (Fig 3.1, Fig 3.2).

This is consistent with some previous studies which demonstrated increased Na+

uptake after acid load (Patrick et al., 1997). Patrick et al showed active Na+

uptake occurs 3 h post injection (Patrick et al., 1997). The findings of the current

study are consistent with those findings for both populations of Fundulus

heteroclitus. This net Na+ gain was possibly caused by increased activity of

Na+/acid exchangers in the apical membrane of branchial cells. As stated above,

net Na+ and Cl- flux difference is a very practical and useful means for estimating

acid-base balance in vivo (Patrick and Wood, 1999). The amount of net Na+

uptake corresponds to that of net acid secretion. The acid secretion in Fundulus

heteroclitus is likely to be complete within 6 h post HCl injection (Fig 3.1, 3.2).

Interestingly, this active Na+ uptake occurred much earlier as the experiments

progressed. At 7 d time point, this active Na+ uptake occurred 2 h after HCl

injection on both populations of Fundulus heteroclitus. This is possibly due to the

increased expression of this Na+/acid exchangers caused by previous HCl

injections. All previous studies were conducted over shorter durations (Evans et

Page 87: Osmoregulation and acid-base tolerance in fish

77

al., 2005; Patrick et al., 1997; Wood and Marshall, 1994), thus this study is the

first to demonstrate this adaptive process of acid tolerance in fish.

However significant differences were not detected in this adaptive process

between the two populations of Fundulus heteroclitus (Fig 3.1, 3.2). Other

studies have shown that intraspecific differences do exist among Fundulus

heteroclitus populations (Able and Palmer, 1988a; Scott et al., 2004b), but it is

not likely the case for acid excretion or Na+ uptake.

Fundulus majalis is a marine fish and can tolerate FW when salinity is

gradually decreased. After the acid load, no increase of Na+ uptake was detected

even 6 hours after the injection (Fig 3.3). There was a significant Na+ loss, which

was possibly due to increased permeability to Na+ at paracellular pathways after

acid load. These data indicated that Fundulus majalis failed to survive during

acidic challenge because of its inability to take up Na+ through a Na+/H+

exchange system. Their failure to regulate permeability to Na+ at paracellular

pathways during acidic challenge may also contribute to their death. There is no

published data on acid tolerance of Fundulus majalis until now. These data are

the first to illustrate the physiological processes of Na+ transport in Fundulus

majalis during acidic challenge.

Previous work has demonstrated the importance of claudin proteins in

epithelial paracellular pathway regulations (Koval, 2006). This study selectively

investigated mRNA expressional levels between two populations of Fundulus

heteroclitus post HCl injection. Our data show that claudin 3-like protein mRNA

levels decreased post HCl injection (Fig 3.4A, Fig 3.5A). Tipsmark et al found

Page 88: Osmoregulation and acid-base tolerance in fish

78

claudin 3 protein levels were higher in tilapia in FW than that in SW, which

indicates that claudin 3 may contribute to a „tighter‟ epithelia (Tipsmark et al.,

2008a). For Fundulus heteroclitus undergoing HCl exposure, an increased

paracellular permeability to Cl- would facilitate the excretion of Cl- and thus reach

a better ability to maintain acid-base balance (Wood and Marshall, 1994). In this

experiment, a „tighter‟ epithelia was not desired because a „leaky‟ paracellular

pathway was needed in order to facilitate Cl- excretion. It is plausible to infer that

the decreased claudin 3-like protein in Fundulus heteroclitus during acidic

challenge may be related to increased permeability to Cl- in paracellular

pathways, which facilitates acid tolerance. These findings together indicate the

roles of claudin 3-like proteins in regulating Cl- transport across paracellular

pathways. Further study is needed for a better understanding of the mechanisms

of this process.

Claudin 5-like protein mRNA expressional levels showed no significant

differences throughout the experiment in both populations of Fundulus

heteroclitus post HCl injection (Fig 3.4B, Fig 3.5B). There is no direct study on

roles of claudin 5 protein in acid-base tolerance in fish. However a recent study

showed claudin-5 contributed to forming a „tight‟ brain barrier in zebrafish

(Abdelilah-Seyfried, 2010). Some studies have indicated that claudin-5

decreased paracellular permeability in vivo (Wen et al., 2004). These findings in

zebrafish and mammal are not in accordance with this study, which indicated a

controversial role of claudin-5 protein in regulating permeability and that different

Page 89: Osmoregulation and acid-base tolerance in fish

79

claudin isoforms may cause distinct functions. The specific roles of claudin 5-like

proteins in Fundulus heteroclitus gills are unclear and need further investigation.

Claudin 23-like protein levels increased on both populations of Fundulus

heteroclitus after HCl injection (Fig 3.4C, Fig 3.5C). To the author‟s knowledge,

there are no known reports regarding claudin-23 function in fish osmoregulation

or paracellular permeability in mammals. The up-regulation of claudin-23 in

osmoregulation and acid-base tolerance in Fundulus species indicated an

important role of this protein though specific contribution of this protein is still not

clear now.

Claudin 28-like protein expression remain unchanged which may indicate it

is not actively involved in acid-base regulation in Fundulus species (Fig 3.4D, Fig

3.5D). However Claudin-28 expression levels decreased in Fundulus grandis

following FW exposure based on the study in Chapter 2. Another study did not

find claudin-28 expression change in Atlantic salmon following salinity transfers

(Tipsmark et al., 2008b). No further information could be obtained from mammals

because it does not exist in mammal. The actual role of this claudin 28-like

protein in acid tolerance in Fundulus heteroclitus needs further investigation.

Page 90: Osmoregulation and acid-base tolerance in fish

80

CHAPTER 4: PROSPECTIVES

Previous studies show that Fundulus species exhibit unexpected

complexity in gene expressions. This unexpected complexity includes divergence

in population genetics, physiological adaptation variation, transcriptional

difference, etc. Comparative analyses among Fundulus species provide a

powerful approach to understand physiological and evolutionary variation. This

powerful tool can be used to elucidate the genetic bases of target gene

expression changes. Similar comparative experiments within Fundulus can be

applied to seek better understandings on underlying transcriptional processes

and regulatory mechanisms responsible for population adaptation to

environmental stressors such as salinity, temperature, pH, pollutant etc.

One of the key characteristics during adaptation to salinity challenges in

euryhaline fish is the morphological changes in branchial epithelia. A series of

factors may contribute to this morphological transformation such as hormones,

phosphorylation on some transmembrane proteins, osmotic sensing receptors on

the cell membrane etc. However a definite description to this question is far from

complete. In addition, signal transductions for cell differentiation and apoptosis

need further investigation. A combination of various approaches such as

morphological, molecular, and gene transcriptional study will contribute to reach

a better understanding of this big picture.

Despite the improved understandings in fish osmoregulation and acid-base

tolerance in the past several decades, most of territories are still not fully charted.

For example, How does osmotic sensing work? What are the specific signal

Page 91: Osmoregulation and acid-base tolerance in fish

81

transduction processes involved in protein expression changes responsible for

solute transport across epithelia? What contributes to interspecific and

intraspecific variation in fish osmoregulation? Experiments of whole animal, in

vitro, molecular biology, morphology, comparative analysis could help us reach a

better understanding in this issue.

Fundulus species which osmoregulatory mechanism is a departure from

the traditional model serve as excellent candidates studying fish osmoregulation.

Failure to locate an active Cl- uptake system in these species makes branchial

paracellular ways assume more important roles in maintaining ion homeostasis in

fish during hypoosmotic challenges. Claudins, the key components regulating

permeability of paracellular pathways, require extensive studies in order to

elucidate the actual functions of specific claudins. Immunohistochemistry staining

to specific claudin proteins, gene knockdown (siRNA, morpholinos, etc), protein

structure analysis will all provide direct valuable information to address this

question.

Participation of claudin proteins is crucial for determining the appropriate

permeability characteristics of FW gill epithelia. Unfortunately there is not much

information available on regulatory mechanisms of these proteins. Some studies

find cortisol, prolactin and growth hormone are the upstream regulators of claudin

proteins. However it is hard to explain some controversial findings. For example,

cortisol exhibits increased expression during both hypo- and hyperosmotic

challenges. The general nature of cortisol effect turning into a specific

development in either direction remains unknown. Interactions between cortisol

Page 92: Osmoregulation and acid-base tolerance in fish

82

and other endocrine or paracrine factors or with the osmotic strain itself need

further investigation.

The linkage between osmoregulation and acid-base tolerance in fish is far

from complete too. Previous studies show extensive interspecific and

intraspecific variation in osmoregulation and acid-base tolerance among fish

species. Do these variations come from differential gene expression? Or from the

availability of some specific genes? How much effect do ecological environments

exert on these variations? Obviously comparative analyses study will help us

elucidate these questions.

Page 93: Osmoregulation and acid-base tolerance in fish

83

REFERENCES CITED

Abdelilah-Seyfried, S., 2010. Claudin-5a in developing zebrafish brain barriers: another brick in the wall. BioEssays 32.

Able, K.W., Palmer, R.E., 1988a. Salinity effects of Fertilization success and Larval Mortality of Fundulus heteroclitus. Copia, 345-350.

Adams, S.M., Lindmeier, J.B., Duvernell, D.D., 2006. Microsatellite analysis of the phylogeography, Pleistocene history and secondary contact hypotheses for the killifish, Fundulus heteroclitus. Mol Ecol 15, 1109-1123.

Alexandre, M.D., Jeansonne, B.G., Renegar, R.H., Tatum, R., Chen, Y.H., 2007. The first extracellular domain of claudin-7 affects paracellular Cl- permeability. Biochem. Biophys. Res. Comm. 357, 87-91.

Alexandre, M.D., Lu, Q., Chen, Y.H., 2005. Overexpression of claudin-7 decreases the paracellular Cl– conductance and increases the paracellular Na+ conductance in LLC-PK1 cells. Journal of Cell Science 118, 2683-2693.

Angelow, S., Yu, A.S.L., 2007. Claudins and paracellular transport: an update. Current Opinion in Nephrology and Hypertension 16, 459-464.

Avella, M., Bornancin, M., 1989. A new analysis of ammonia and sodium transport through the gills of the freshwater rainbow trout (Salmo gairdneri. 142, 155-175.

Bagherie-Lachidan, M., Wright, S.I., Kelly, S.P., 2008. Claudin-3 tight junction proteins in Tetraodon nigroviridis: cloning, tissue-specific expression, and a role in hydromineral balance. Am. J. Physiol. 294, R1638-R1647.

Bagherie-Lachidan, M., Wright, S.I., Kelly, S.P., 2009. Claudin-8 and-27 tight junction proteins in puffer fish Tetraodon nigroviridis acclimated to freshwater and seawater. J. Comp. Physiol. B. 179, 419-431.

Bernardi, G., Sordino, P., Powers, D.A., 1993. Concordant Mitochondrial and Nuclear-DNA Phylogenies for Populations of the Teleost Fish Fundulus heteroclitus. Proc. Nat. Acad. Sci. 90, 9271-9274.

Page 94: Osmoregulation and acid-base tolerance in fish

84

Biosen, A., Amstrup, J., Novak, I., Grosell, M., 2003. Sodium and chloride transport in soft water and hard water acclimated zebrafish (Danio rerio). Biochimica et Biophysica Acta (BBA) - Bioenergetics 1618, 207-218.

Burnett, K.G., Bain, L.J., Baldwin, W.S., Callard, G.V., Cohen, S., Di Giulio, R.T., Evans, D.H., Gomez-Chiarri, M., Hahn, M.E., Hoover, C.A., Karchner, S.I., Katoh, F., MacLatchy, D.L., Marshall, W.S., Meyer, J.N., Nacci, D.E., Oleksiak, M.F., Rees, B.B., Singer, T.D., Stegeman, J.J., Towle, D.W., Van Veld, P.A., Vogelbein, W.K., Whitehead, A., Winn, R.N., Crawford, D.L., 2007a. Fundulus as the premier teleost model in environmental biology: Opportunities for new insights using genomics. Comparative Biochemistry and Physiology D-Genomics & Proteomics 2, 257-286.

Cameron, J.N., 1978. Regulation of blood pH in teleost fish. Respiratory Physiology 33, 129-144.

Cameron, J.N., 1984. Acid-Base status of fish at different temperatures. Am J Physiol 246, R452-R459.

Chang, I.C., Hwang, P.P., 2004. Cl- uptake mechanism in freshwater-adapted tilapia (Oreochromis mossambicus). Physiol. Biochem. Zool. 77, 406-414.

Chang, I.C., Lee, T.H., Yang, C.H., Wei, Y.W., Chou, F.I., Hwang, P.P., 2001. Morphology and function of gill mitochondria-rich cells in fish acclimated to different environments. Physiol. Biochem. Zool. 74, 111-119.

Chiba, H., Osanai, M., Murata, M., Kojima, T., Sawada, N., 2008. Transmembrane proteins of tight junctions. Biochimica Et Biophysica Acta-Biomembranes 1778, 588-600.

Chidester, F.E., 1920. The Behavior of Fundulus heteroclitus on the Salt Marshes of New Jersey. The American Naturalist, 551-557.

Claiborne, J.B., Perry, E., Bellows, S., Campbell, J., 1997. Mechanisms of acid-base excretion across the gills of a marine fish. J. Exp. Zool. 279, 509-520.

Clelland, E.S., Bui, P., Bagherie-Lachidan, M., Kelly, S.P., 2010. Spatial and salinity-induced alterations in claudin-3 isoform mRNA along the

Page 95: Osmoregulation and acid-base tolerance in fish

85

gastrointestinal tract of the pufferfish Tetraodon nigroviridis. Comp. Biochem. Physiol. A. 155, 154-163.

Copeland, D.E., 1950. Adaptive behavior of the chloride cell in the gill of Fundulus heteroclitus. Journal of Morphology 87, 369-379.

Coyne, C.B., Gambling, T.M., Boucher, R.C., Carson, J.L., Johnson, L.G., 2003a. Role of claudin interactions in airway tight junctional permeability. lung cellular and molecular physiology, 1166-1178.

Duggins, C.F., Relyea, K.G., Karlin, A.A., 1989. Biochemical systematics in southeastern populations of Fundulus heteroclitus and Fundulus grandis. Northeast Gulf Science. 10, 95-102.

Duranti, E., Ehrenfeld, J., Harvey, B.J., 1986. Acid secretion through the Rana esculenta skin: involvement of an anion-exchange mechanism at the basolateral membrane. The Journal of Physiology.

Duvernell, D.D., Lindmeier, J.B., Faust, K.E., Whitehead, A., 2008. Relative influences of historical and contemporary forces shaping the distribution of genetic variation in the Atlantic killifish, Fundulus heteroclitus. Mol Ecol 17, 1344-1360.

Edwards, S.L., Toop, T., 2002. Immunolocalisation of NHE3-like immunoreactivity in the gills of the rainbow trout (Oncorhynchus mykiss) and the blue-throated wrasse (Pseudolabrus tetrious). Journal of Anatomy 195, 465-469.

Esaki, M., Hoshijima, K., Kobayashi, S., Fukuda, H., Kawakami, K., Hirose, S., 2007. Visualization in zebrafish larvae of Na+ uptake in mitochondria-rich cells whose differentiation is dependent on foxi3a. Am. J. Physiol. 292, R470-R480.

Evans, D.H., 2008. Teleost fish osmoregulation: what have we learned since August Krogh, Homer Smith, and Ancel Keys. American Journal of Physiological Regulatory Integretive Comparative Physiology 295, R704-713.

Evans, D.H., 2010. A brief history of the study of fish osmoregulation: the central role of the Mt. Desert Island Biological Laboratory. Frontiers in Aquatic Physiology 1.

Page 96: Osmoregulation and acid-base tolerance in fish

86

Evans, D.H., Claiborne, J.B., 2006. The Physiology of Fishes. Book.

Evans, D.H., Piermarini, P.M., Choe, K.P., 2005. The multifunctional fish gill: Dominant site of gas exchange, osmoregulation, acid-base regulation, and excretion of nitrogenous waste. Physiol. Rev. 85, 97-177.

Evans, D.H., Piermarini, P.M., Potts, W.T.W., 1999. Ionic transport in the fish gill epithelium. J. Exp. Zool. 283, 641-652.

Fangue, N.A., Hofmeister, M., Schulte, P.M., 2006. Intraspecific variation in thermal tolerance and heat shock protein gene expression in common killifish, Fundulus heteroclitus. J. Exp. Biol. 209, 2859-2872.

Feldmeth, C.R., Waggoner, J.P., 1972. Field Measurements of Tolerance to Extreme Hypersalinity in the California Killifish, Fundulus parvipinnis. Copeia 3, 592-594.

Fenwick, J.C., Wendelaar Bonga, S.E., Flik, G., 1999. In vivo bafilomycin-sensitive Na(+) uptake in young freshwater fish. J. Exp. Biol. 202 Pt 24, 3659-3666.

Fromm, M., Schulzke, J.D., 2009. Molecular structure and function of the tight junction. . Annals of the New York Academy of Sciences 1165.

Fumi, K., Susumu, H., Toyoji, K., 2002. Vacuolar-type proton pump in the basolateral plasma membrane energizes ion uptake in branchial mitochondria-rich cells of killifish Fundulus heteroclitus, adapted to a low ion environment. The Journal of Experimental Biology 206, 793-803.

Galvez, F., Franklin, N.M., Tuttle, R.B., Wood, C.M., 2007. Interactions of waterborne and dietary cadmium on the expression of calcium transporters in the gills of rainbow trout: influence of dietary calcium supplementation. Aquat. Toxicol. 84, 208-214.

Galvez, F., Reid, S.D., Hawkings, G., Goss, G.G., 2001. Isolation and characterization of mitochondria-rich cell types from the gill of freshwater rainbow trout. Am J Physiol 282, R658-R668.

Page 97: Osmoregulation and acid-base tolerance in fish

87

Gary, V.S., Steve, F.P., James, N.F., 1996. Localization of mRNA for the proton pump (H+-ATPase) and exchanger in the rainbow trout gill. Canadian Journal of Zoology 74, 2095-2103.

George, W.K., Christopher, W.P., Robert, L.P., 2006a. Energetics of osmoregulation: I. oxygen consumption by Fundulus heteroclitus. The Journal of Experimental Zoology 305, 309-317.

Gilmour, K.M., Bayaa, M., Kenney, L., McNeill, B., Perry, S.F., 2006. Type IV carbonic anhydrase is present in the gills of spiny dogfish (Squalus acanthias). Comparative and Evolutionary Physiology 292, 556-567.

Goggisberg, A., Hesse, M., 1983. Putrescine, spermidine, spermine, and related polyamine alkaloids. Alkaloids 22, 85-188.

Gonzalez-villasenor, L., Powers, D.A., 1990. Mitochondrial-DNA Restriction-site Polymorphisms in the Teleost Fundulus heteroclitus Support Secondary Intergradation. Evolution 44.

Goss, G.G., Adamia, S., Galvez, F., 2001. Peanut lectin binds to a subpopulation of mitochondria-rich cells in the rainbow trout gill epithelium. Am J Physiol 281, R1718-R1725.

Goss, G.G., Perry, S.F., Fryer, J.N., Laurent, P., 1998. Gill morphology and acid-base regulation in freshwater fishes, 107-115.

Goss, G.G., Perry, S.F., Wood, C.M., Laurent, P., 1992. Mechanisms of ion and acid-base regulation at the gills of freshwater fish. J. Exp. Zool. 263, 143-159.

Goss, G.G., Wood, C.M., 1990a. Na+ and Cl- Uptake Kinetics, Diffusive Effluxes and Acidic Equivalent Fluxes across the Gills of Rainbow-Trout .1. Responses to Environmental Hyperoxia. J. Exp. Biol. 152, 521-547.

Goss, G.G., Wood, C.M., 1990b. Na+ and Cl- uptake kinetics, diffusive effluxes and acidic equivalent fluxes across the gills of rainbow trout. 2. Responses to bicarbonate infusion. J. Exp. Biol. 152, 549-571.

Page 98: Osmoregulation and acid-base tolerance in fish

88

Griffith, R.W., 1974a. Environment and salinity tolerance in the Genus Fundulus. Copeia, 319-331.

Haag, W.R., Warren, M.L., 2000. Effects of light and presence of fish on lure display and larval release behaviours in two species of freshwater mussels. Animal Behaviour 60, 879-886.

Heisler, N., 1984. Acid-base regulation in fishes. Fish Physiology XA, 315-401.

Hirata, T., Kaneko, T., Ono, T., Nakazato, T., Furukawa, N., Hasegawa, S., Wakabayashi, S., Shigekawa, M., Chang, M.H., Romero, M.F., Hirose, S., 2003. Mechanism of acid adaptation of a fish living in a pH 3.5 lake. Am. J. Physiol. 284, R1199-R1212.

Hiroaki, M., Toyoji, K., Shinichi, U., Sei, S., Yoshio, T., 2002. Kidney-specific chloride channel, OmClC-K, predominantly expressed in the diluting segment of freshwater-adapted tilapia kidney. Proc. Nat. Acad. Sci. 99, 15782-15787.

Hiroi, J., McCormick, S.D., Ohtani-Kaneko, R., Kaneko, T., 2005. Functional classification of mitochondrion-rich cells in euryhaline Mozambique tilapia (Oreochromis mossambicus) embryos, by means of triple immunofluorescence staining for Na/K+-ATPase, Na+/K+/2Cl(-) cotransporter and CFTR anion channel. J. Exp. Biol. 208, 2023-2036.

Hoffmann, E.K., 1992. Cell swelling and volume regulation. 70 Suppl, S310-S313.

Hossler, F.E., Musil, G., Karnaky, K.J.J., Epstein, F.H., 1985. Surface ultrastructure of the gill arch of the killifish, Fundulus heteroclitus, from seawater and freshwater, with special reference to the morphology of apical crypts of chloride cells. 185, 377-386.

Hsiao, S., Meier, A.H., 2005. Comparison of Semilunar Cycles of Spawning Activity in Fundulus grandis and Fundulus heteroclitus held under constant laboratory conditions. Camparative Physiology and Biochemistry 252, 213-218.

Page 99: Osmoregulation and acid-base tolerance in fish

89

Hwang, P.-P., Lee, T.-H., 2007. New insights into fish ion regulation and mitochondrion-rich cells. Comparative Biochemistry and Physiology - Part A 148, 479-497.

Hyde, D.A., Perry, S.F., 1987. Acid-base and ionic regulation in the American eel ( Anguilla rostrata ) during and after prolonged aerial exposure: branchial and renal adustments. J. Exp. Biol. 133, 429-447.

Inokuchi, M., Hiroi, J., Watanabe, S., Lee, K.M., Kaneko, T., 2008. Gene expression and morphological localization of NHE3, NCC and NKCC1a in branchial mitochondria-rich cells of Mozambique tilapia (Oreochromis mossambicus) acclimated to a wide range of salinities. Comp Biochem Physiol A 151, 151-158.

Kalujnaia, S., 2007. Transcriptomic approach to the study of osmoregulation in the European eel Anguilla anguilla. Physiological Genomics 31, 385-401.

Kaneko, T., Katoh, F., 2004. Functional morphology of chloride cells in killifish Fundulus heteroclitus, a euryhaline teleost with seawater preference. Fish. Sci. 70, 723-733.

Karnaky, K.J., 1980. Ion-Secreting Epithelia - Chloride Cells in the Head Region of Fundulus-Heteroclitus. Am J Physiol 238, R185-R198.

Karnaky, K.J., 1986. Structure and function of the chloride cell of Fundulus heteroclitus and other teleosts. American Zoologist 26, 209-224.

Karnaky, K.J.J., Degnan, K.J., Garretson, L.T., Zadunaisky, J.A., 1984. Identification and quantification of mitochondria-rich cells in transporting epithelia. 246, R770-R775.

Karnaky, K.J.J., Kinter, L.B., Kinter, W.B., Stirling, C.E., 1976. Teleost chloride cell. II. Autoradiographic localization of gill Na,K-ATPase in killifish Fundulus heteroclitus adapted to low and high salinity environments. 70, 157-177.

Katoh, F., Hasegawa, S., Kita, J., Takagi, Y., Kaneko, T., 2001. Distinct seawater and freshwater types of chloride cells in killifish, Fundulus heteroclitus. Can. J. Zool. 79, 822-829.

Page 100: Osmoregulation and acid-base tolerance in fish

90

Katoh, F., Kaneko, T., 2003. Short-term transformation and long-term replacement of branchial chloride cells in killifish transferred from seawater to freshwater, revealed by morphofunctional observations and a newly established 'time-differential double fluorescent staining' technique. J. Exp. Biol. 206, 4113-4123.

Kirschner, L.B., 1997. Extrarenal mechanisms in hydromineral and acid-base regulation in aquatic vertebrates. In: The Handbook of Physiology., 577-622.

Koval, M., 2006. Claudins - Key pieces in the tight junction puzzle. Cell Communication and Adhesion 13, 127-138.

Krogh, A., 1938. The active transport of ions in some freshwater animals. 25, 335-350.

Krogh, A., 1939. Osmotic regulation in aquatic animals. . Originally published by Cambridge University Press and reprinted in English in 1965 in an unabridged and unaltered edition, 242 pp.

Kumai, Y., Bahubeshi, A., Steele, S., Perry, S.F., 2011. Strategies for maintaining Na+ balance in zebrafish (Danio rerio) during prolonged exposure to acidic water. Comparative Biochemistry and Physiology - Part A 160, 52-62.

Laurent, P., Chevalier, C., Wood, C.M., 2006. Appearance of cuboidal cells in relation to salinity in gills of Fundulus heteroclitus, a species exhibiting branchial Na+ but not Cl- uptake in freshwater. Cell and Tissue Research 325, 481-492.

Laurent, P., Dunel, S., 1980. Morphology of gill epithelia in fish. 238, R147-R159.

Lin, C.H., Huang, C.L., Yang, C.H., Lee, T.H., Hwang, P.P., 2004. Time-course changes in the expression of Na, K-ATPase and the morphometry of mitochondrion-rich cells in gills of euryhaline tilapia (Oreochromis mossambicus) during freshwater acclimation. J Exp Zool 301A, 85-96.

Lin, H., Randall, D.J., 1991. Evidence for the presence of an electrogenic proton pump on the trout gill epithelium. J. Exp. Biol. 161, 119-134.

Page 101: Osmoregulation and acid-base tolerance in fish

91

Loeb, J., Wasteneys, H., 1912. On the adaptation of fish (Fundulus) to higher temperatures. J. Exp. Zool. 12, 543-557.

Loh, Y.H., Christoffels, A., Brenner, S., Hunziker, W., Venkatesh, B., 2004. Extensive expansion of the claudin gene family in the teleost fish, Fugu rubripes. Genome Res 14, 1248-1257.

Maetz, J., 1976. Transport of ions and water across the epithelium of fish gills. 133-159.

Marshall, W.S., 2003. Rapid regulation of NaCl secretion by estuarine teleost fish: coping strategies for short-duration freshwater exposures. Biochimica et Biophysica Acta (BBA) - Biomembranes 1618, 95-105.

Marshall, W.S., Bryson, S.E., Luby, T., 2000. Control of epithelial Cl- secretion by basolateral osmolality in the euryhaline teleost Fundulus heteroclitus. J. Exp. Biol. 203, 1897-1905.

Marshall, W.S., Emberley, T.R., Singer, T.D., Bryson, S.E., McCormick, S.D., 1999. Time course of salinity adaptation in a strongly euryhaline estuarine teleost, Fundulus heteroclitus: A multivariable approach. J. Exp. Biol. 202, 1535-1544.

Marshall, W.S., Grosell, M., 2005. Ion transport, osmoregulation, and acid-base balance, in: D.H. Evans, J.B. Claiborne (Eds.), Physiology of Fishes. CRC Press, Boca Raton, FL, 177-230.

Marshall, W.S., Lynch, E.A., Cozzi, R.R.F., 2002. Redistribution of immunofluorescence of CFTR anion channel and NKCC cotransporter in chloride cells during adaptation of the killifish Fundulus heteroclitus to sea water. J. Exp. Biol. 205, 1265-1273.

Marshall, W.S., Ossum, C.G., Hoffmann, E.K., 2005. Hypotonic shock mediation by p38 MAPK, JNK, PKC, FAK, OSR1 and SPAK in osmosensing chloride secreting cells of killifish opercular epithelium. J. Exp. Biol. 208, 1063-1077.

Marvao, P., Emilio, M.G., Gil Ferreira, K.G., Fernandes, P.L., Gil Ferreira, H., 1994. Ion transport in the intestine of Anguilla anguilla : gradients and translocators. J. Exp. Biol. 193, 97-117.

Page 102: Osmoregulation and acid-base tolerance in fish

92

McCormick, S.D., Dickhoff, W.W., Duston, J., Nishioka, R.S., Bern, H.A., 1991. Developmental differences in the responsiveness of gill Na+,K(+)-ATPase to cortisol in salmonids. 84, 308-317.

McCormick, S.D., Sundell, K., Björnsson, B.T., Brown, C.L., Hiroi, J., 2003. Influence of salinity on the localization of Na+/K+-ATPase, Na+/K+/2Cl- cotransporter (NKCC) and CFTR anion channel in chloride cells of the Hawaiian goby (Stenogobius hawaiiensis). The Journal of Experimental Biology 206, 4575-4583.

Morita, K., Furuse, M., Fujimoto, K., Tsukita, S., 1999. Claudin multigene family encoding four-transmembrane domain protein components of tight junction strands. Proc. Nat. Acad. Sci. 96, 511-516.

Parry, G., Holliday, F., Blaxter, J., 1959. Chloride secretory cells in the gills of teleosts. Nat. 183, 1248-1249.

Patrick, M.L., Part, P., Marshall, W.S., Wood, C.M., 1997. Characterization of ion and acid-base transport in the fresh water adapted mummichog (Fundulus heteroclitus). J. Exp. Zool. 279, 208-219.

Patrick, M.L., Wood, C.M., 1999. Ion and acid-base regulation in the freshwater mummichog (Fundulus heteroclitus): a departure from the standard model for freshwater teleosts. Comparative Biochemistry and Physiology a-Molecular and Integrative Physiology 122, 445-456.

Perry, S.F., Furimsky, M., Bayaa, M., Georgalis, T., Shahsavarani, A., Nickerson, J.G., Moon, T.W., 2003a. Integrated responses of Na+/HCO3

- cotransporters and V-type H+-ATPases in the fish gill and kidney during respiratory acidosis. Biochim Biophys Acta 1618, 175-184.

Perry, S.F., Shahsavarani, A., Georgalis, T., Bayaa, M., Furimsky, M., Thomas, S.L., 2003b. Channels, pumps, and exchangers in the gill and kidney of freshwater fishes: their role in ionic and acid-base regulation. Journal of Experimental Zoology Part A 300, 53-62.

Philpott, C.W., Copeland, D.E., 1963. Fine structure of chloride cells from three species of Fundulus Journal of Cell Biology 18, 389-404.

Primer3, http://frodo.wi.mit.edu/primer3/.

Page 103: Osmoregulation and acid-base tolerance in fish

93

Salama, A., Morgan, I.J., Wood, C.M., 1999. The linkage between Na+ uptake and ammonia excretion in rainbow trout: kinetic analysis, the effects of (NH4)2SO4 and NH4HCO3 infusion and the influence of gill boundary layer pH. Journal of Experimental Biology 202, 697-709.

Sas, D., Hu, M.C., Moe, O.W., Baum, M., 2008. Effect of claudins 6 and 9 on paracellular permeability in MDCK II cells. Am. J. Physiol. 295, R1713-R1719.

Scott, G.R., Claiborne, J.B., Edwards, S.L., Schulte, P.M., Wood, C.M., 2005. Gene expression after freshwater transfer in gills and opercular epithelia of killifish: insight into divergent mechanisms of ion transport. J. Exp. Biol. 208, 2719-2729.

Scott, G.R., Richards, G.J., Forbush, B., Isenring, P., Schulte, P.M., 2004a. Changes in gene expression in gills of the euryhaline killifish Fundulus heteroclitus after abrupt salinity transfer. American Journal of Physiology- Cell Physiology 287, 300-309.

Scott, G.R., Rogers, J.T., Richards, J.G., Wood, C.A., Schulte, P.M., 2004b. Intraspecific divergence of ionoregulatory physiology in the euryhaline teleost Fundulus heteroclitus: possible mechanisms of freshwater adaptation. J. Exp. Biol. 207, 3399-3410.

Scott, G.R., Schulte, P.M., 2005. Intraspecific variation in gene expression after seawater transfer in gills of the euryhaline killifish Fundulus heteroclitus. Comp Biochem Physiol A 141, 176-182.

Shehadeh, Z.H., Gordon, M.S., 1969. Role of Intestine in Salinity Adaptation of Rainbow Trout Salmo Gairdneri. Comp Biochem Physiol 30, 397-&.

Shmukler, B.E., Kurschat, C.E., Ackermann, G.E., Jiang, L., Zhou, Y., Barut, B., Stuart-Tilley, A.K., Zhao, J., Zon, L.I., Drummond, I.A., Vandorpe, D.H., Paw, B.H., Alper, S.L., 2005. Zebrafish slc4a2/ae2 anion exchanger: cDNA cloning, mapping, functional characterization, and localization. Am J Physiol Renal Physiol 289, F835-849.

Singer, T.D., Tucker, S.J., Marshall, W.S., Higgins, C.F., 1998. A divergent CFTR homologue: highly regulated salt transport in the euryhaline teleost F. heteroclitus. American Journal of Physiology 274, C715-723.

Page 104: Osmoregulation and acid-base tolerance in fish

94

Smith, M.W., Chapman, R.W., Powers, D.A., 1998. Mitochondrial DNA analysis of Atlantic Coast, Chesapeake Bay, and Delaware Bay populations of the teleost Fundulus heteroclitus indicates temporally unstable distributions over geologic time. Molecular Marine Biology and Biotechnology 7.

Stewart, P.A., 1983. Modern quantitative acid-base chemistry. Canadian Journal of Physiology and Pharmacology 61, 1444-1461.

Sullivan, G.V., Fryer, J.N., Perry, S.F., 1995. Immunolocalization of proton pumps (H+-ATPase) in pavement cells in rainbow trout gill. J. Exp. Biol. 198, 2619-2629.

Sumner, F.B., 1911. Fundulus and fresh water. Sci. 34, 928-931.

Tang, C.H., Lee, T.H., 2007. The effect of environmental salinity on the protein expression of Na+/K+-ATPase, Na+/K+/2Cl- cotransporter, cystic fibrosis transmembrane conductance regulator, anion exchanger 1, and chloride channel 3 in gills of a euryhaline teleost, Tetraodon nigroviridis. Comparative Biochemistry and Physiology - Part A 147, 521-528.

Tang, V.W., Goodenough, D.A., 2003. Paracellular ion channel at the tight junction. Biophysical Journal 84, 1660-1673.

Tipsmark, C.K., Baltzegar, D.A., Ozden, O., Grubb, B.J., Borski, R.J., 2008a. Salinity regulates claudin mRNA and protein expression in the teleost gill. Am J Physiol 294, R1004-R1014.

Tipsmark, C.K., Kiilerich, P., Nilsen, T.O., Ebbesson, L.O.E., Stefansson, S.O., Madsen, S.S., 2008b. Branchial expression patterns of claudin isoforms in Atlantic salmon during seawater acclimation and smoltification. American Journal of Physiological Regulatory Integrative Comparative Physiology 294, R1563-1574.

Tipsmark, C.K., Madsen, S.S., Borski, R.J., 2004. Effect of salinity on expression of branchial ion transporters in striped bass (Morone saxatilis). Journal of Experimantal Zoology 301A, 979-991.

Tipsmark, C.K., Madsen, S.S., Seidelin, M., Christensen, A.S., Cutler, C.P., Cramb, G., 2002. Dynamics of Na(+),K(+),2Cl(-) cotransporter and Na(+),K(+)-ATPase expression in the branchial epithelium of brown trout (Salmo

Page 105: Osmoregulation and acid-base tolerance in fish

95

trutta) and Atlantic salmon (Salmo salar). Journal of Experimantal Zoology 293, 106-118.

Tipsmark, C.K., Sorensen, K.J., Madsen, S.S., 2010. Aquaporin expression dynamics in osmoregulatory tissues of Atlantic salmon during smoltification and seawater acclimation. J. Exp. Biol. 213, 368-379.

Tresguerres, M., Katoh, F., Fenton, H., Jasinska, E., Goss, G.G., 2005. Regulation of branchial V-H+-ATPase Na+/K+-ATPase and NHE2 in response to acid and base infusions in the Pacific spiny dogfish (Squalus acanthias). J. Exp. Biol. 208, 345-354.

Van Itallie, C.M., Anderson, J.M., 2004. The molecular physiology of tight junction pores. Physiology 19, 331-338.

Van Itallie, C.M., Anderson, J.M., 2006. Claudins and epithelial paracellular transport. Ann Rev Physiol 68, 403-429.

Vernier, P.J.B.a.J.-M., 1989. Plasma lipoproteins in fish. Journal of Lipid Research 30, 467-489.

Wen, H.J., Watry, D.D., Cecilia, M., Marcondes, G., Fox, H.S., 2004. Selective Decrease in Paracellular Conductance of Tight Junctions: Role of the First Extracellular Domain of Claudin-5. Mol. Cell. Biol. 24, 8408-8417.

Whitehead, A., 2009. Comparative mitochondrial genomics within and among species of killifish. BMC Evolutionary Biology 9.

Whitehead, A., 2010. The evolutionary radiation of diverse osmotolerant physiologies in killifish (Fundulus sp.). Evolution 64, 2070-2085.

Whitehead, A., Crawford, D.L., 2006. Variation within and among species in gene expression: raw material for evolution. Mol Ecol 15, 1197-1211.

Whitehead, A., Galvez, F., Zhang, S., Williams, L.M., Oleksiak, M.F., 2010. Functional Genomics of Physiological Plasticity and Local Adaptation in Killifish. Journal of Heredity.

Page 106: Osmoregulation and acid-base tolerance in fish

96

Whitehead, A., Roach, J.L., Zhang, S.J., Galvez, F., 2011. Genomic mechanisms of evolved physiological plasticity in killifish distributed along an environmental salinity gradient. Proceedings of the Academy of Natural Sciences of North America 108, 6193-6198.

Whitehead A, T.D., Champlin D, Nacci D, 2010. Comparative transcriptomics implicates mechanisms of evolved pollution tolerance in a killifish population. Mol Ecol 19, 5186-5203.

Wilson, J.M., Laurent, P., 2002. Fish gill morphology: inside out. Journal of Experimental Zoology 293, 192-213.

Wilson, J.M., Laurent, P., Tufts, B.L., Benos, D.J., Donowitz, M., Vogl, A.W., Randall, D.J., 2000. NaCl uptake by the branchial epithelium in freshwater teleost fish: an immunological approach to ion-transport protein localization. J. Exp. Biol. 203 Pt 15, 2279-2296.

Wilson, J.M., Whiteley, N.M., Randall, D.J., 2002. Ionoregulatory changes in the gill epithelia of coho salmon during seawater acclimation. Physiol. Biochem. Zool. 75, 237-249.

Wilson, R.W., WOOD, C.M., Gonzalez, R.J., Patrick, M.L., Bergman, H.L., Narahara, A., Val, A.L., 1995. Ion and acid-base balance in three species of Amazonian fish during gradual acidification of extremely soft water. 72, 277-285.

Wood, C.A., Laurent, P., 2003. Na+ versus Cl- transport in the intact killifish after rapid salinity transfer. Biochimica Et Biophysica Acta-Biomembranes 1618, 106-119.

Wood, C.M., Bucking, C., Grosell, M., 2010. Acid-base responses to feeding and intestinal Cl- uptake in freshwater- and seawater-acclimated killifish, Fundulus heteroclitus, an agastric euryhaline teleost. J. Exp. Biol. 213, 2681-2692.

Wood, C.M., Gilmour, K.M., Part, P., 1998. Passive and active transport properties of a gill model, the cultured branchial epithelium of the freshwater rainbow trout (Oncorhynchus mykiss). Comparative Biochemistry Physiology A 119, 87-96.

Page 107: Osmoregulation and acid-base tolerance in fish

97

Wood, C.M., Grosell, M., 2008. A critical analysis of transepithelial potential in intact killifish (Fundulus heteroclitus) subjected to acute and chronic changes in salinity. J. Comp. Physiol. B. 178, 713-727.

Wood, C.M., Grosell, M., 2009. TEP on the tide in killifish (Fundulus heteroclitus): effects of progressively changing salinity and prior acclimation to intermediate or cycling salinity. J. Comp. Physiol. B. 179, 459-467.

Wood, C.M., LeMoigne, J., 2001. Intracellular acid-base responses to environmental hyperoxia and normoxic recovery in rainbow trout. Respiration Physiology 86, 91-113.

Wood, C.M., Marshall, W.S., 1994. Ion balance, acid-base regulation, and chloride cell function in the common killifish, Fundulus heteroclitus - a euryhaline estuarine teleost. Estuaries 17, 34-52.

Wood, C.M., Wheatly, M.G., Hobe, H., 1984. The mechanisms of acid-base and ionoregulation in the freshwater rainbow trout during environmental hyperoxia and subsequent normoxia. III. Branchial exchanges. Respiration Physiology 55, 175-192.

Zall, D.M., Fisher, D., Garner, M.Q., 1956. Photometric Determination of Chlorides in Water. Anal. Chem. 28, 1665-1668.

Page 108: Osmoregulation and acid-base tolerance in fish

98

VITA

Shujun Zhang is the oldest son among the three of Weidong Zhang and

Xiaofeng Wang. He graduated with a Bachelor of Medicine degree from Tongji

Medical College in Wuhan, China, in July of 2003. Then he started doing his

residency in Liver Cancer in Tongji Hospital and obtained his Master degree from

Huazhong University of Science&Technology in 2005. He entered an orthopedic

residency program in Wuhan Orthopedics Hospital in August of 2005. He started

his doctoral work in Louisiana State University in Baton Rouge, Louisiana from

fall of 2006. Mr. Zhang will graduate with the degree of Doctor of Philosophy in

Biological Sciences in the Summer of 2012.