17.01.2015 Views

Kenneth Halberg

Kenneth Halberg

Kenneth Halberg

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Osmoregulation | With Focus on Fluid<br />

and Solute Dynamics in Tardigrada<br />

PhD Dissertation<br />

KENNETH A. HALBERG<br />

© <strong>Kenneth</strong> A. <strong>Halberg</strong>


FACULTY OF SCIENCE<br />

UNIVERSITY OF COPENHAGEN<br />

Osmoregulation│With Focus on Fluid<br />

and Solute Dynamics in Tardigrada<br />

PhD Dissertation<br />

<strong>Kenneth</strong> A. <strong>Halberg</strong><br />

UNIVERSITAS<br />

HAFNIENSIS<br />

2012<br />

Dissertation submitted Monday the 14th of May 2012.<br />

Supervisor: Associate Professor Nadja Møbjerg, PhD.


Dissertation presented at University of Copenhagen to be publicly examined (provided acceptance in its<br />

current form) in Auditorium 1, August Krogh Building, Universitetsparken 13, Thursday, June 28, 2012<br />

at 14:00 for the degree of Doctor of Philosophy. The examination will be conducted in English.<br />

Abstract<br />

<strong>Halberg</strong>, K. A. 2012. Osmoregulation │With Focus on Fluid and Solute Dynamics in<br />

Tardigrada.<br />

Osmoregulation is the regulated control of water and solute composition in body fluid<br />

compartments. On one hand, the internal composition must be kept within optimal<br />

conditions for metabolic processes in the face of external perturbation. On the other<br />

hand, the nature of the living state demands a continuous traffic of compounds in and<br />

out of the organism. These demands appear to be in fundamental contradiction<br />

however cells and animals achieve so-called “steady-state” by means of an array of<br />

transport proteins, which provide a stringent control on the exchange of water and<br />

solutes across body surfaces. The distinct mechanisms of solute transport have been<br />

studied in most animal groups, but there are still large gaps in our understanding of<br />

how animals cope with osmotic stress. In the present thesis, osmoregulatory<br />

phenomena were studied in vertebrate and invertebrate organism alike, with the main<br />

focus being on fluid and solute dynamics in Tardigrada. For example, the inorganic<br />

ion composition of several species was investigated, which revealed that tardigrades<br />

contain roughly similar relative contributions of inorganic ions to total osmotic<br />

concentration, when compared to closely related animal groups. Moreover, it was<br />

inferred that cryptobiotic tardigrades (species able to enter a state of latent life)<br />

contain a large fraction of organic osmolytes. The mechanisms of organic anion<br />

transport in a marine species of tardigrade was investigated pharmacologically, and<br />

compared to that of insects. These data showed that organic anion transport is<br />

localized to the midgut epithelium and that the transport is both active and transporter<br />

mediated with a pharmacological profile similar to that of insects. Tardigrades survive<br />

in a variety of osmotic environments (semi-terrestrial, limnic and marine habitats),<br />

why the ability to volume and osmoregulate was examined. These studies demonstrated<br />

an ability to regulate total body volume during both hypo- and hyperosmotic<br />

conditions, and that the ability to hyper-regulate could be a general theme among<br />

members of eutardigrades. Thus, the work presented herein, have contributed to<br />

establishing tardigrades as an important experimental group in which central<br />

physiological questions may be answered, including aspects of osmotic and ionic<br />

regulation.<br />

Keywords: osmoregulation, volume regulation, organic anion transport, hyper-regulate,<br />

inorganic ions, organic osmolytes, tardigrade, insect,<br />

<strong>Kenneth</strong> A. <strong>Halberg</strong>, The August Krogh Centre, Department of Biology,<br />

Universitetsparken 13, DK-2100 Copenhagen Ø, Denmark<br />

© <strong>Kenneth</strong> A. <strong>Halberg</strong> 2012


“Beautiful is what we see,<br />

More beautiful is what we perceive,<br />

Most beautiful is what we do not understand”<br />

- Niels Stensen


List of Papers<br />

This thesis is based on the following papers and manuscripts, which are referred to<br />

in the text by their Roman numerals.<br />

I. <strong>Halberg</strong>, K. A., Larsen, K. W., Jørgensen, A., Ramløv, H. & Møbjerg,<br />

N. Cryptobiotic tardigrades contain large fraction of unidentified<br />

organic solutes: A comparative study on inorganic ion composition in<br />

Tardigrada.<br />

II.<br />

<strong>Halberg</strong>, K. A., & Møbjerg, N. (2012). First evidence of epithelial<br />

transport in tardigrades: A comparative investigation of organic anion<br />

transport. Journal of Experimental Biology, 215: 497-507.<br />

III. Møbjerg, N. M., <strong>Halberg</strong>, K. A., Jørgensen, A. Persson., D., Bjørn M,<br />

Ramløv H & Kristensen R. M. (2011). Survival in extreme<br />

environments – on current knowledge of adaptations in tardigrades.<br />

Acta Physiologica, 202: 409-420.<br />

IV. Haugen, B.M., <strong>Halberg</strong>, K.A., Jespersen, Å., Prehn, L.R. & Møbjerg,<br />

N. (2010). Functional characterization of the vertebrate primary ureter:<br />

Structure and ion transport mechanisms of the pronephric duct of<br />

axolotl larvae (Amphibia). BMC developmental Biology, 10: 56.<br />

V. <strong>Halberg</strong>, K. A., Persson, D., Ramløv, H., Westh, P., Kristensen, R. M.<br />

& Møbjerg, N. (2009). Cyclomorphosis in Tardigrada: Adaption to<br />

environmental constraints. Journal of Experimental Biology, 212:<br />

2803-2811.<br />

Additionally, the following papers and manuscripts were prepared during the<br />

course of my PhD studies, but are not included in the thesis:<br />

VI.<br />

VII.<br />

<strong>Halberg</strong> K. A., Jørgensen, A. and Møbjerg, N. (in prep.). Surviving<br />

without water: Tun formation in tardigrades is an active process<br />

mediated by the musculature<br />

<strong>Halberg</strong> K. A., Persson, D., Jørgensen, A. Kristensen, R. M. and<br />

Møbjerg, N. (submitted). Population dynamics of a marine tardigrade:<br />

Temperature limits geographic distribution of Halobiotus crispae.<br />

Marine Biological Research


VIII. Persson, D., <strong>Halberg</strong> K. A., Jørgensen A., Møbjerg N. & Kristensen<br />

R. M. (in review). Neuroanatomy of Halobiotus crispae (Eutardigrada:<br />

Hypsibiidae): Tardigrade brain structure suggests inclusion into<br />

Panarthropoda. Journal of Morphology.<br />

IX. Persson, D., <strong>Halberg</strong> K. A., Jørgensen A., Ricci C., Møbjerg N. &<br />

Kristensen R. M. (2010). Extreme stress tolerance in tardigrades:<br />

Surviving space conditions in low earth orbit. Journal of Zoological<br />

Systematics and Evolutionary Research, 49: 90-97.<br />

X. <strong>Halberg</strong>, K. A., Persson D., Møbjerg N., Wanninger A. & Kristensen<br />

R. M. (2009). Myoanatomy of the Marine Tardigrade Halobiotus<br />

crispae (Eutardigrada: Hypsibiidae). Journal of Morphology, 270:<br />

996-1013.<br />

Lastly, following paper provides important background knowledge for the work<br />

presented herein:<br />

XI. Møbjerg, N., A. Jørgensen, J. Eibye-Jacobsen, K. A. <strong>Halberg</strong>, D.<br />

Persson & R. M. Kristensen (2007). New Records on cyclomorphosis<br />

in the marine eutardigrade Halobiotus crispae (Eutardigrada:<br />

Hypsibiidae). Journal of Limnology, 66 (suppl. 1): 132-140.


Reprint and publication is made with permission from the respective copyright<br />

holders.<br />

Paper II, V © The Company of Biologists.<br />

Paper IV, is copyright of the authors.<br />

Paper III © Wiley-Blackwell<br />

Statement of authorship<br />

Paper I: KAH was deeply involved in study design. KAH participated in<br />

extracting animals and ion chromatography. KAH performed nanoliter osmometry.<br />

KAH performed the data analysis, prepared the figures, participated in discussions<br />

and interpretation of the data, and drafted the manuscript.<br />

Paper II: KAH performed the major part of the experimental work and data<br />

analysis. He participated in planning of experiments, data interpretation, prepared<br />

the figures and drafted the manuscript.<br />

Paper III: KAH performed cell counts and provided images of tardigrades. He<br />

helped draft parts of the manuscript.<br />

Paper IV: KAH participated in immunostaining experiments, performed CLSM<br />

and prepared the 3D images. KAH participated in discussions and interpretation of<br />

the data.<br />

Paper V: KAH participated in planning of experiments, sampling, staging,<br />

scanning electron microscopy, DSC experiments, experiments on cold hardiness<br />

and osmotic stress tolerance, volume measurements, hemolymph sample<br />

collections, and nanoliter osmometry. He furthermore participated in discussions<br />

and interpretation of data, prepared the figures and drafted the manuscript.<br />

Front cover: Scanning elecron micrographs of the tardigrades Rictersius coronifer<br />

(top left), Halobiotus crispae (middle right), and Echiniscus testudo (middle<br />

bottom).


Contents<br />

Preface.......................................................................................................................... 9<br />

Introduction ................................................................................................................ 11<br />

Maintaining a stable internal environment .......................................................... 11<br />

Osmoregulators and osmoconformers ................................................................. 12<br />

Osmoregulatory organs........................................................................................ 12<br />

Filtration-Reabsorption systems ................................................................ 13<br />

Secretion-Reabsorption systems ................................................................ 14<br />

Phylum Tardigrada..................................................................................................... 17<br />

General morphology ............................................................................................ 18<br />

Classification ....................................................................................................... 19<br />

Ecology ................................................................................................................ 21<br />

Fluid and solute dynamics – an overview .................................................................. 23<br />

Inorganic ion composition ................................................................................... 23<br />

Organic anion transport ....................................................................................... 25<br />

Volume and osmoregulation................................................................................ 26<br />

Conclusions and future perspectives .......................................................................... 27<br />

Dansk sammenfatning ................................................................................................ 29<br />

Acknowledgements .................................................................................................... 30<br />

References .................................................................................................................. 32


Preface<br />

The primary aim of this thesis was to address several aspects of the fluid and<br />

solute dynamics in tardigrades, and hereby provide new insight into the general<br />

stress biology of these enigmatic creatures. This was done by adopting an<br />

integrative approach, i.e. applying advanced methods in biology, analytic<br />

biochemistry and physical chemistry, which offered functional data from different<br />

disciplines to be obtained. The experimental work was carried out mainly at The<br />

August Krogh Centre, University of Copenhagen; however, additional<br />

experimental work was performed at the Danish Natural History Museum,<br />

University of Copenhagen and at the Department of Nature, Systems and Models,<br />

Roskilde University. Overall this thesis has contributed to converting tardigrades,<br />

from an almost exclusive taxonomic phenomenon into an established and<br />

important experimental group, in which central physiological questions can be<br />

investigated.<br />

This PhD-thesis comprises a short introduction to osmotic and ionic<br />

regulation in Metazoa, accompanied by a brief review on the general morphology,<br />

classification and ecology of tardigrades. Moreover, an overview of the results<br />

presented as well as conclusions and future perspectives are presented. Five papers<br />

and manuscripts form the basis of this thesis, of which four are published in peer<br />

review journals (Papers II, III, IV and V), and one is prepared for submission<br />

(Paper I). I am the first author of three (Paper I, II and V) and second author on<br />

two (Papers III and IV) of these papers. Five additional papers and manuscripts<br />

were prepared during the course of my PhD studies, but are not included in the<br />

thesis. This work was funded by the 2008 Faculty of Science, University of<br />

Copenhagen Freja-Programme.<br />

Copenhagen, the 14 th of May 2012<br />

<strong>Kenneth</strong> A. <strong>Halberg</strong><br />

9


Introduction<br />

Maintaining a stable internal environment<br />

“Life is as a thing of macromolecular cohesion in salty water” (Gilles & Delpire,<br />

1997). Albeit a crude statement, it frames the fact that the ability to control salt<br />

and water balance is a fundamental prerequisite for both cellular and animal life.<br />

Indeed, the internal environment must usually be kept within relatively narrow<br />

limits, as substantial deviations in cell composition are incompatible with the<br />

optimal function of macromolecules (lipids, proteins, RNA), and may ultimately<br />

modify the rate and extent of cellular reactions (Zhao, 2005). The overall<br />

mechanism by which animals conserve a proper osmotic balance between cells,<br />

extracellular fluid and the environment is termed osmoregulation.<br />

The basis for osmoregulation lies in the strict control of the ionic<br />

composition and the osmotic pressure of the intra- and extracellular compartments<br />

through the regulated accumulation and loss of inorganic ions and organic<br />

compounds (Dawson & Liu, 2009). This regulation is achieved through the<br />

coordinated activity of an array of transporter proteins (both energy-consuming<br />

and passive), which collectively maintain the steady-state condition of cells and<br />

animals (Essig, 1968). During steady-state conditions, compositions of the intraand<br />

extracellular compartments are maintained in a non-equilibrium state (Dawson<br />

& Liu, 2009). This uneven distribution of solutes is important for keeping an<br />

optimal milieu for metabolic processes (Zhao, 2005). Accordingly, the<br />

extracellular fraction of the body fluids of animals are typically high in Na + and Cl -<br />

, and relatively low in the other major ions (e.g. K + , Ca 2+ and Mg 2+ ), while the<br />

intracellular environment of most organisms is low in Na + but high in K + , PO 4<br />

3-<br />

and proteins (e.g. Dawson & Lui, 2009; Paper I). As such, the plasma membrane<br />

of cells must maintain ionic, but not osmotic, differences, while specialized<br />

excretory organs – e.g. antennal glands of crustaceans, Malpighian tubules of<br />

insects and tardigrades, rectal glands of sharks and rays, gills and intestine of<br />

teleost fishes, salt glands of birds and reptiles, the kidneys of vertebrates etc. –<br />

often maintain both ionic and osmotic differences between animals and their<br />

environments (Riegel, 1970; Peaker, 1971; Paper IV, Beyenbach & Piermarini,<br />

2011; Reilly et al., 2011; Whittamore, 2012). In general, mechanisms that allowed<br />

organisms to respond and adapt to an osmotic challenge over the course of<br />

11


evolution, has been fundamental to the invasion of new osmotic hostile habitats,<br />

and such ecological divergence in turn is an important mechanism for the<br />

speciation process (Schluter, 2009). Accordingly, if the ability to osmoregulate<br />

had not evolved, life on the planet would look quiet different from how we know<br />

it.<br />

Osmoregulators and osmoconformers<br />

Generally, animals are divided into two broad categories in terms of their<br />

responses to osmotic stress: osmoregulators, which maintain an internal osmolarity<br />

different from that of the external environment, and osmoconformers, which<br />

conform to the external medium in which they are immersed (Fig. 1). Most<br />

vertebrates are strict osmoregulators (e.g. Paper IV), as they maintain ionic and<br />

osmotic balance within narrow limits; although hagfish, a basal group of<br />

vertebrates, are a notable exception (Sardella et al., 2009). Conversely, marine<br />

invertebrates are typically categorized as osmoconformers, as many of them<br />

appear to be in osmotic balance with sea water over a range of salinities (Fig. 1).<br />

However, there are numerous obvious exceptions including the marine tardigrade<br />

Halobiotus crispae as well as several members of Crustacea (e.g. Sarver et al.,<br />

1994; Normant et al., 2005; Paper V). In fact, the terms ‘strict osmoregulator’ and<br />

‘strict osmoconformer’ must be used with caution, as typical osmoregulators are<br />

forced to conform during the most extreme conditions (e.g. Dowd et al., 2010),<br />

whereas animals otherwise described as osmoconformers actually maintain slight<br />

differences between the internal and external environment (e.g. van Weel, 1957).<br />

Although achieved through different mechanisms, both osmoregulating and<br />

osmoconforming animals may tolerate wide ranges of external salinities, thus<br />

termed euryhaline species, while animals intolerant of large changes are called<br />

stenohaline species.<br />

Osmoregulatory organs<br />

Osmoregulatory organs are specialized organs involved in maintaining ionic and<br />

osmotic homeostasis in the face of osmotic perturbation, as well as in excreting<br />

endobiotic and exobiotic waste products (Dawson & Lui, 2009; Papers II, IV).<br />

The specific organs mediating these processes may vary between different groups<br />

of animals (see above); however, the molecular basis and specific mechanism of<br />

solute and water transport often show a highly convergent/homologous pattern<br />

across different types of epithelia (e.g. Paper II). In general, two types of systems<br />

12


Fig. 1 Examples of osmotic performance of representative species from various groups<br />

exposed to sea/brackish water, expressed as the relation between internal (extracellular) and<br />

external (habitat) osmolality. These data show that osmoconformaty (∆osm=0) is present in<br />

invertebrate and vertebrate species alike, albeit strong hypo-regulators and the ability to<br />

produce a hyperosmotic urine appears restricted to vertebrates. It should be emphasized that<br />

the selected species not necessarily represent the osmotic performance of the entire group, as<br />

large difference may exist between even closely related species. For original data see:<br />

Robertson, 1949 a ; van Weel, 1957 e ; Dice, 1968 d ; Liggins and Grigg, 1985 j ; Diehl, 1986 b ;<br />

Normant et al., 2005 f ; Sardella et al., 2009 g ; Paper V c ;Reilly et al., 2011 h ; Whittamore,<br />

2012 i ).<br />

have evolved by which the initial process of urine formation takes place i) the<br />

filtration-reabsorption type and ii) the secretion-reabsorption type.<br />

Filtration-Reabsorption systems<br />

The kidneys of vertebrates (fish, amphibians, reptiles, birds and mammals), and<br />

the functional analogs of crusteans and molluscs, maintain extracellular fluid<br />

homeostasis by producing urine through the filtration of plasma (ultrafiltration),<br />

which is subsequently modified by selective reabsorption and secretion of ions,<br />

organic molecules and water (Anderson, 1960; Schmidt-Nielsen, 1963; Riegel,<br />

13


1970; Møbjerg et al., 2004; Paper IV; Whittamore et al., 2012). In vertebrates,<br />

three temporally and spatially different kidney generations, the pronephroi,<br />

mesonephroi and metanephroi, successively maintain fluid and electrolyte<br />

homeostasis during morphogenesis, with the pronephroi constituting the functional<br />

kidneys of fish and amphibian larvae (Paper IV). The functional unit of the<br />

vertebrate kidney is the nephron, which is composed of a filtration unit and a renal<br />

tubule (Anderson, 1960; Møbjerg et al., 2004; Paper IV). The filtration process<br />

that takes place in the filtration unit (the glomerulus and Bowman’s capsule) is<br />

‘passive’ i.e. entirely driven by the hydrostatic pressure generated by the heart,<br />

whereas the reabsorption and secretion processes take place over specialized<br />

epithelia of the renal tubule (Schmidt-Nielsen, 1963). The primary membrane<br />

transporter for energizing vertebrate tissue is the Na/K-ATPase (e.g. Paper IV).<br />

Accordingly, vertebrate kidneys may produce both dilute, iso-osmotic and<br />

concentrated urine relative to the body fluids (Paper IV; Whittamore et al., 2012),<br />

which has been a dominant factor in allowing vertebrates to penetrate into all types<br />

of habitats on Earth. Filtration-reabsorption systems are capable of processing<br />

large volumes of fluids, but are energetically costly, as any substance (e.g.<br />

glucose) that has been filtered remains in the urine unless subsequently reabsorbed<br />

(Schmidt-Nielsen, 1963). The advantage of such a system; however, is that new<br />

potentially toxic compounds are eliminated without the need for developing<br />

distinct secretory pathways for each new compound, which may be necessary for<br />

the secretion-reabsorption type system (Paper II).<br />

Secretion-Reabsorption systems<br />

The Malpighian tubules of insects, and possibly tardigrades (Møbjerg & Dahl,<br />

1996; Papers II, III, V), are the functional analogs of the vertebrate kidney, but<br />

constitute a secretion-reabsorption system that produces urine in a fundamentally<br />

different way than the filtration-reabsorption-systems (Beyenbach & Piermarini,<br />

2011). In the absence of blood vessels (i.e. a closed circulatory system), the<br />

hemolymph of insects is circulated at pressures insufficient for filtration, and the<br />

Malpighian tubules thus form the (primary) urine entirely by secretion (Beyenbach<br />

& Piermarini, 2011). The formation of the primary urine is generally initiated in<br />

the distal segments (blind-ended tip) of the Malpighian tubule, and is essentially<br />

iso-osmotic (consisting mainly of KCl and NaCl) to the hemolymph (Williams &<br />

Beyenbach, 1983). The subsequent reabsorption of water, ions and metabolites in<br />

proportions that maintain extracellular homeostasis (a process analogous to that of<br />

vertebrates) takes place in downstream structures i.e. proximal tubule, hindgut and<br />

rectum (O’Donnell & Maddrell, 1995; Coast, 2007). The final urine composition is<br />

14


adjusted in the rectum (Coast, 2007) and may be either strongly hypo- or<br />

hyperosmotic depending on the species and its physiological status (Maddrell &<br />

Phillips, 1975; Reynolds & Bellward, 1989). In contrast to vertebrate epithelia, the<br />

V-type H + -ATPase is considered ubiquitous in energizing insect epithelia<br />

(Beyenbach & Piermarini, 2011); although the Na/K-ATPase is still expressed and<br />

functionally relevant for tubular function (Torrie et al., 2004; Paper II). In fact,<br />

energized by the H + -ATPase, some of the highest fluid secretion rates per unit area<br />

membrane from any tissue have been reported from hematophagous insects (e.g.<br />

Aedes aegypti, Rhodnius prolixus) after a blood meal (Williams & Beyenbach,<br />

1983; Maddrell & Phillips, 1975). In addition to playing a key role in<br />

osmoregulation, new properties of the Malpighian tubules of insects have emerged<br />

in recent years, which suggest that Malpighian tubules are involved in such diverse<br />

functions as renal detoxification, metabolism of toxins and immune system<br />

responses (Dow & Davies, 2006; Paper II).<br />

As holds for insects, ultrastructural studies on the Malpighian tubules of<br />

tardigrades indicate that they function as secretion-reabsorption systems involved<br />

in fluid and solute transport (Weglarska, 1987; Møbjerg and Dahl, 1996; Peltzer et<br />

al., 2007). They are positioned at the transition zone between the midgut and<br />

rectum of eutardigrades (Fig. 2), and the positional conformity between insects<br />

and eutardigrades has been used as a strong argument for their homology (Greven,<br />

1982; Møbjerg and Dahl, 1996). However, at present no functional data exist<br />

relating the Malpighian tubules of tardigrades to an osmoregulatory role.<br />

Accordingly, functional studies on the fluid and solute dynamics of tardigrades are<br />

greatly needed (Papers I, II, III, V), and due to the close affinity to the<br />

euarthropod complex (Aguinaldo et al., 1997), would be highly useful in<br />

understanding and reconstructing the evolution of osmoregulation in Insects and<br />

other arthropods.<br />

15


Fig. 2 Structure and organization of the Malpighian tubules of the marine eutardigrade<br />

Halobiotus crispae. dm, dorsal Malpighian tubule; dp, distal part; is, initial segment; mg,<br />

midgut; mv, microvilli,; nu, nucleus; pp, proximal part; re, rectum; From: Møbjerg & Dahl,<br />

1996.<br />

16


Phylum Tardigrada<br />

Tardigrades, also known as water bears, are among the smallest multi-cellular<br />

animals on the planet (0.1-1.2 mm). They were discovered in 1773 by the German<br />

pastor J. A. E. Goeze, who described them as “kleiner Wasserbär”, or little water<br />

bear, due to their strong resemblance to a tiny bear (Ramazzotti & Maucci 1982).<br />

Not long after the current name “Tardigrada” was given by the Italian naturalist<br />

Spallanzani in 1776 (Lat. tardus – slow, grado – walker). In response to<br />

unfavorable environmental conditions, many species of tardigrades have the ability<br />

to enter the ametabolic state of suspended animation, also known as cryptobiosis,<br />

in which the organism is neither dead nor alive (Møbjerg et al., 2011; Fig. 3). The<br />

animal can remain in this state for as much as 20 years (Jørgensen et al., 2007), yet<br />

once external conditions again become favorable, the tardigrade resumes activity<br />

unaffected. This incredible ability is shared with selected species of nematodes,<br />

rotifers and arthropods (Glasheen & Hand, 1988; Crowe & Maddin, 1974; Ricci et<br />

al., 2003). In 1962, Tardigrada was recognized as a phylum by Ramazzotti in Il<br />

Phylum Tardigrada. Presently, there are more than 1000 described species<br />

(Guidetti & Bertolani 2005; Degma & Guidetti, 2007; Degma et al., 2012);<br />

however, it has been estimated that several thousand species remain undescribed<br />

(Paper III).<br />

Tardigrada is included in the invertebrate superclade Ecdysozoa<br />

(Aguinaldo et al. 1997); however, their precise phylogenetic position is still being<br />

debated. Both molecular and morphological investigations produce conflicting<br />

conclusions, and it is currently unclear whether the group is more closely related to<br />

the nematodes and nematomorphs or to arthropods and onychophorans (Aguinaldo<br />

et al. 1997; Dunn et al. 2008; Zantke et al., 2008; Edgecombe 2010; Rota-Stabelli<br />

et al., 2010; Cambell et al., 2011). Regardless, this group is closely related to one<br />

of the two most species-rich and economically important groups Nematoda or<br />

Arthropoda, and thus maintains a central position in relation to the two major<br />

invertebrate model organisms, i.e. Caenorhabditis elegans Maupas, 1900 and<br />

Drosophila (Sophophora) melanogaster Meigen, 1830 (Gabriel et al. 2007;<br />

Goldstein and Blaxter 2002).<br />

17


Fig. 3 Scanning electron micrographs showing the external morphology of Richtersius<br />

coronifer (Eutardigrada) A. lateral view of the active, hydrated state B. Dorsal view of the<br />

dehydrated, cryptobiotic state C. Ventral view of the dehydrated, cryptobiotic state. The<br />

pictures illustrate the morphological changes associated with entry into an ametabolic state<br />

(i.e. cryptobiosis), which include the retraction of head and limbs into the body cavity, and the<br />

formation of a compact shape – the tun. From: Paper VI.<br />

General morphology<br />

Tardigrades are bilaterally symmetric micrometazoans with a body divided into<br />

five separate body segments, i.e. a cephalic segment, containing a mouth, eyespots<br />

18


and sensory organs (papillae cephalica or cirri and clavae), and four trunk<br />

segments (Nelson, 2002; Fig. 3A). The first three trunk segments each bear a pair<br />

of lateroventrally directed legs, while the terminal trunk segment bears a pair of<br />

posterioventrally directed legs (Figs. 3, 4). The legs typically terminate in 4 to 13<br />

claws or suction discs (Nelson, 2002). Tardigrades are ventrally flattened with a<br />

convex dorsal side, and are covered by a segmented cutinous cuticle, which is<br />

periodically shed during molting – formation of the new cuticle is maintained by a<br />

single layer of epidermal cells (Nelson, 2002). The digestive system consists of a<br />

foregut, midgut and a hindgut with a pair of stylets and stylet glands flanking the<br />

buccal tube. Three glands (the Malpighian tubules) are positioned at the transition<br />

zone between the midgut and hindgut in eutardigrades (Weglarska, 1987; Møbjerg<br />

and Dahl, 1996; Peltzer et al., 2007). Tardigrades posses a hemocoel-type of fluidfilled<br />

body cavity, i.e. an open circulatory system as seen in arthropods and<br />

nematodes, which likely functions in circulation and respiration. The somatic<br />

musculature is composed of structurally independent muscle fibers, which can be<br />

divided into a dorsal, ventral, dorsoventral, and a lateral musculature in addition to<br />

a distinct leg musculature (Schmidt-Rhaesa & Kulessa, 2007; Fig. 4). Moreover,<br />

the buccopharyngeal muscles, intestinal muscles and cloacal muscles comprise the<br />

animal’s visceral musculature. Whereas cross striation of the somatic musculature<br />

is especially pronounced in Arthrotardigrada, the somatic muscles of Eutardigrada<br />

are described as an intermediate between smooth and obliquely striated (Walz,<br />

1974). The nervous system of tardigrades consists of an (at least) three lobed brain<br />

(Fig. 5) and a ventral nerve cord with four fused paired ganglia that shows a clear<br />

segmental organization.<br />

Classification<br />

Originally based on morphological characters, tardigrades are divided into two<br />

main evolutionary lines, represented by the extant lineages Eutardigrada and<br />

Heterotardigrada. The validity of a third class, Mesotardigrada, is currently<br />

considered dubious (Ramazzotti and Maucci, 1983).<br />

Heterotardigrada consists of the orders Arthrotardigrada and Echiniscoidea<br />

with arthrotardigrades possessing the most plesiomorphic characters.<br />

Arthrotardigrada consists exclusively of marine species (Renaud-Mornant 1982;<br />

Jørgensen et al. 2010), and are morphologically the most diverse group. They are<br />

present in all oceans from intertidal zones to abyssal depths, and inhabit various<br />

types of sediment. In contrast, the Echiniscoidea comprises both limno-terrestrial,<br />

limnic as well as marine species with the majority of the described species<br />

belonging to the family Echiniscidae. Taxonomically, the main characters<br />

19


Fig. 4 Tardigrade musculature as revealed by fluorescently coupled phalloidin in combination<br />

with confocal laser scanning microscopy three-dimensional reconstruction A. Lateral view of<br />

Halobiotus crispae (Eutardigrada), Paper VIII B. Ventral view of Echiniscoides sigismundi<br />

(Heterotardigrada), unpublished data.<br />

separating the two groups from Eutardigrada include a separate gonopore, a closed<br />

three-lobed anus as well as well-developed cephalic-, trunk-, and leg appendages<br />

(Guidetti and Bertolani, 2005).<br />

The eutardigrades are divided into the two orders; Apochela and Parachela.<br />

Both orders contain mainly limno-terrestrial species, albeit with a few exceptions –<br />

20


Fig. 5 Conceptual drawing based on immunofluorescent and ultrastructural data, showing an<br />

interpretation of the brain structure of Halobiotus crispae. A. Lateral view B. Frontal view.<br />

clg, claw gland; co, connective; dc, dorsal commissure; ey, eye; g0, sub-pharyngeal ganglion;<br />

gI, first ventral trunk ganglion; ic, inner connective; il, inner lobe; lgg, leg ganglion; mg,<br />

median ganglion; mo, mouth opening; oc, outer connective; ol, outer lobe; pc, papilla<br />

cephalica; pb, pharyngeal bulb; st, stylet; t, temporalia; vll, ventrolateral lobe. From: Paper<br />

VIII.<br />

most notably the marine genus Halobiotus (see Paper III). In general,<br />

eutardigrades are cylindrically shaped with a more or less distinct segmentation,<br />

and exhibit a relatively uniform morphology (Fig. 3A). The key characters of<br />

eutardigrades include a cloaca (combined gonopore and anus), the presence of<br />

Malpighian tubules and a strong reduction of cephalic sensory structures (Guidetti<br />

and Bertolani, 2005). Structures such as the bucco-pharyngeal apparatus and claw<br />

shape are important taxonomic characters within Eutardigrada.<br />

Ecology<br />

Tardigrades occupy a range of moisture regimes and often constitute a major<br />

component of meiofaunal communities in terrestrial, limnic and marine<br />

ecosystems throughout the globe. However, they are distinctly aquatic organisms,<br />

requiring a film of water to be active. Tardigrades are predominantly egg-laying,<br />

with both sexual and parthenogenetic modes of reproduction described (Bertolani,<br />

2001). Molting occurs continuously throughout their lifecycle, which may be<br />

between 3-30 months (Nelson, 2002). Populations of tardigrades have been<br />

studied in a variety of habitats; including mosses, lichens, leaf litter and soil, and<br />

21


Fig. 6 Population dynamics of the marine tardigrade Halobiotus crispae showing a unimodal<br />

pattern of maximal frequency. Graphic representation of sampling data (2006-2012)<br />

comparing the temporal pattern in abundance of H. crispae to abiotic parameters<br />

(temperature, (─ ─); salinity, (- - -); pH, (- ─ -), and the seasonal appearance of the different<br />

cyclomorphic stages (shown on top), from the locality of Vellerup Vig, Isefjord, Denmark.<br />

Light grey area is the period in which exuvia containing eggs were found, and thus indicates<br />

the period of sexual reproduction. From: Paper VII.<br />

the life history and population dynamics have received some attention (Martinez,<br />

1975; Morgan 1977; Guidetti et al. 1999; Uhía and Briones 2002; Jönson 2003;<br />

Suzuki 2003). Tardigrade population dynamics may show both unimodal and<br />

bimodal patterns of annual variation (Martinez, 1975; Morgan, 1977; Fig. 6),<br />

albeit the specific pattern appears to be both species and habitat specific. Factors<br />

such as temperature, moisture and food availability have been suggested to be<br />

correlated with population density (Hallas & Yeates, 1972; Morgan, 1977).<br />

However, other factors including competition, predation and parasitism may play a<br />

role in controlling population density and diversity (Nelson, 2002). Tardigrades<br />

possess an amazing reproductive capacity, as indicated by the large changes in<br />

animal density on short temporal scales (Morgan, 1977).<br />

22


Fluid and solute dynamics – an overview<br />

Inorganic ion composition<br />

Knowledge of the composition as well as concentrations of dissolved particles in<br />

internal fluids of an organism, and how these change during various exposures, is<br />

fundamental to the understanding of its basic physiology. However, practically<br />

nothing is known about these aspects in tardigrades, which has been a major<br />

obstacle to the study of the fluid and solute dynamics in these animals. In Paper I,<br />

the inorganic ion content of five different species (Echiniscus testudo, Milnesium<br />

tardigradum, Richtersius coronifer, Macrobiotus cf. hufelandi and Halobiotus<br />

crispae) covering both a large phylogenetic and habitat spectrum was analyzed<br />

(Fig. 7A). These data demonstrated that Na + and Cl - are the principle inorganic<br />

ions in tardigrade fluids, albeit substantial concentrations of K + , NH + 4 , Ca 2+ ,<br />

Mg 2+ , F - 2-<br />

3-<br />

, SO 4 and PO 4 were also detected. Moreover, tardigrades appear to<br />

contain roughly similar relative contributions of the respective inorganic ions to<br />

total osmotic concentration, when compared to selected species of arthropods,<br />

nematodes and onychophorans (Hobson et al., 1952; Sutcliffe, 1962; Campiglia,<br />

1975; Wilder et al., 1998; Normant et al., 2005), albeit a large relative contribution<br />

of calcium appears characteristic of tardigrade fluids. Inorganic ions destabilize<br />

macromolecules and affect the rate and extent of metabolic reactions at high<br />

concentrations, which invariably leads to impairment of cellular function (Zhao,<br />

2005; Yancey, 2005). However, apparently R. coronifer does not exclude<br />

inorganic ions during dehydration (Fig. 7B), which suggests a concomitant<br />

accumulation of organic solutes. Moreover, as evidenced by the differences<br />

between the calculated osmotic concentrations of the known ions and the<br />

measured total osmolarity in the different species (Fig. 7A), it was inferred that<br />

cryptobiotic tardigrades (in steady-state) contain a large fraction of unidentified<br />

organic osmolytes. Organic osmolytes can be divided into a few major categories<br />

(sugars, polyols, amino acids and various derivatives), and several of these groups<br />

possess known protective functions in relation to osmotic stress (Hincha and<br />

Hagemann, 2004; Yancey, 2005). Accordingly, the future detection and analysis of<br />

such compounds are likely to provide new insight into the biochemistry and<br />

physiology of superior tardigrade adaptations.<br />

23


Fig. 7 Ionic contributions to total osmotic concentration A. Concentrations (mM) of the<br />

respective cations and anions measured in each investigated species, as well as the<br />

corresponding total osmotic concentration (mOsm/kg). The blank area represents the osmotic<br />

deficit (OD). B. Concentrations (mg/l) of the respective cations and anions measured in<br />

hydrated, active specimens compared to dehydrated cryptobiotic animals of Richtersius<br />

coronifer. Data are expressed as mean ± s.d. From: Paper I.<br />

24


Organic anion transport<br />

The ability to excrete endogenous<br />

waste products as well as environmental<br />

toxins is an essential step to<br />

avoid these compounds reach toxic<br />

levels and to keep metabolic reactions<br />

within optimal conditions. One of the<br />

better known systems involved in such<br />

an excretion is the organic anion<br />

transport system, the function of<br />

which has been studied in several<br />

vertebrate (fish, amphibians, reptiles,<br />

birds and mammals) and invertebrate<br />

(nematodes and insects) model<br />

organisms (George et al., 1999;<br />

Dantzler, 2002; Dow and Davies,<br />

2006). In Paper II, these data were<br />

expanded as the sites, characteristics<br />

and pharmacological profile of the<br />

transepithelial transport of chlorophenol<br />

red (CPR), a prototypical<br />

Fig. 8 Tentative cellular model for the<br />

transepithelial transport of organic anions in<br />

tardigrades. From: Paper II.<br />

substrate of the classic organic anion secretion pathway, was investigated in the<br />

tardigrade Halobiotus crispae Kristensen, 1982 and compared to corresponding<br />

data from the desert locust Schistocerca gregaria Forskål, 1775. This was done by<br />

introducing a new method for quantifying non-fluorescent dyes. Our study<br />

revealed i) that tardigrades posses an organic anion transport system, ii) that it was<br />

localized to the midgut epithelium, and iii) that organic anion secretion was both<br />

active and transporter mediated, with possible members of the SLC21/SLCO<br />

transport families mediating the basolateral entry step in tardigrade midgut cells.<br />

Transport by insect Malpighian tubules showed a similar pharmacological profile,<br />

but higher concentrations of CPR were achieved. Based on the observed transport<br />

characteristics in the presence and absence of transport inhibitors, a tentative<br />

cellular model for the transepithelial transport of CPR in tardigrades was<br />

suggested (Fig. 8). Specifically, a large lumen positive potential generated by the<br />

H + -ATPase could provide the driving force for accumulation of anions in the<br />

lumen, although the exact coupling between electrochemical gradients generated<br />

by the pumps and transport of ions is unknown. This study was the first to provide<br />

evidence for epithelial transport in tardigrades.<br />

25


Volume and osmoregulation<br />

Tardigrades survive in a variety of osmotic environments and must protect the<br />

internal tissues from the vagaries and extremes of the external environment. The<br />

secondary marine species Halobiotus crispae colonizes habitats characterized by<br />

especially large fluctuations in external salinity, why the ability to respond to an<br />

osmotic challenge was investigated in this species (Paper V). Animals were<br />

exposed to both hypo- and hyperosmotic media and the subsequent changes in<br />

total body volume and internal osmotic pressure were recorded. These data<br />

revealed that H. crispae is able to regulate total body volume back to control<br />

values when immersed in hypotonic solutions, yet was unable to do so in<br />

concentrated media. Instead a new steady-state was achieved significantly below<br />

control conditions. Animal activity was only markedly affected in very dilute<br />

media, suggesting a possible effect on neuro-muscular function at low salt<br />

concentrations. Conversely, when analyzing the concomitant changes in<br />

hemolymph concentrations, it appeared that H. crispae is a euryhaline<br />

osmoconformer, in which the hemolymph osmotic pressure is largely governed by<br />

the external environment. However, expressing the hemolymph osmolality<br />

measured during exposure to dilute as well as concentrated media as a function of<br />

external salinity revealed that H. crispae is in fact a strong hyper-regulator (Fig.<br />

9A). Expanding these studies to include the limno-terrestrial species Richtersius<br />

coronifer Richters, 1903 (Paper III) showed that this is could be a general feature<br />

of all eutardigrades (Fig. 9B), which was further supported by data presented in<br />

Paper I. The ability to hyper-regulate indicates the excretion of dilute urine. The<br />

three glands positioned at the transition zone between the midgut and rectum of<br />

eutardigrades, are generally considered to have an excretory function. This<br />

assumption is based on the probable homology of the Malpighian tubules of<br />

eutardigrades and insects (Greven, 1982; Møbjerg and Dahl, 1996), and on several<br />

ultrastructural studies that shows that the epithelium is likely involved in fluid and<br />

solute transport (Weglarska, 1987; Møbjerg and Dahl, 1996; Peltzer et al., 2007).<br />

However, data in support of an osmoregulatory function of both the rectum<br />

(Dewel and Dewel, 1979) and midgut (Paper II) also exists, which emphasizes the<br />

need for functional studies on these organs at both the molecular and cellular level.<br />

26


Fig. 9 Osmotic performance of A. Halobiotus crispae and B. Richtersius coronifer during<br />

exposure to media of varying osmotic strength. From: Paper III.<br />

27


Conclusions and future perspectives<br />

This dissertation has provided new insight into the fluid and solute dynamics of<br />

metazoans, particularly relating to one of the most enigmatic groups on the planet<br />

– the tardigrades. Using a multi-disciplinary approach, crucial information was<br />

provided on both organs and systems of several species, representing vertebrates,<br />

arthropods and tardigrades, and general patterns in especially tardigrade<br />

physiology have emerged (e.g. Paper III). By comparing our data on tardigrades<br />

to several evolutionary related groups, including nematodes, onychophorans and<br />

arthropods, basic physiological principles have been discovered (e.g. Paper II),<br />

which emphasizes the importance of comparative physiology.<br />

In this respect, future work on tardigrade physiology could encompass an<br />

extension of the work presented herein. Additional species from different habitats<br />

and evolutionary lineages should be investigated, in order to further explore the<br />

diversity as well as common trends in tardigrade biology. In particular, studies on<br />

marine cryptobionts (e.g. Echiniscoides sigismundi) and additional noncryptobiotic<br />

species would help clarify whether the osmotic deficits observed in<br />

cryptobiotic animals (Fig. 5; Paper I) in fact are related to cryptobiotic ability or<br />

alternatively to habitat preference. Regardless, large scale analyses and<br />

characterization of the organic solutes of cryptobiotic tardigrades should be<br />

performed, which surely would provide an enhanced resolution of several aspects<br />

relating to tardigrade stress responses. Other studies could include a<br />

characterization of the volume and osmoregulatory capacity of heterotardigrade<br />

species. Our results on eutardigrades show a capacity to hyper-regulate over a<br />

broad range of external salinities (Fig. 7); however, do heterotardigrades without<br />

Malpighian tubules posses the same ability True limnic species (e.g. Bertolanius<br />

nebulosus) should also be investigated. In addition to the suggested whole animal<br />

experiments, studies on the molecular and cellular level are needed to fully<br />

understand how fluid and electrolyte homeostasis is achieved in these animals. For<br />

example, electrophysiological investigations using single cell glass microelectrode<br />

impalements on dissected native tissue would help characterize the cellular<br />

transporters involved in e.g. urine formation. Collectively, these studies would be<br />

important to elucidate how tardigrades functionally have solved colonizing every<br />

major type of habitat on the planet, and perhaps be useful in reconstructing how<br />

osmoregulation has evolved in metazoans.<br />

28


Dansk sammenfatning<br />

Osmoregulering er kontrollen af kropsvæskernes sammensætning af vand og<br />

opløste stoffer. Foruden at opfylde de optimale betingelser for metaboliske<br />

processer under indflydelse af eksterne påvirkninger, skal denne<br />

sammensætning samtidig imødekomme den kontinuerlige transport af stoffer<br />

ind og ud af den levende organisme. Disse krav forekommer umiddelbart<br />

fundamentalt uforenelige, men celler og dyr opnår et såkaldt ”steady-state” ved<br />

hjælp af et spektrum af transportproteiner, der udøver streng kontrol over<br />

udvekslingen af vand og opløste stoffer over forskellige kropsoverflader. De<br />

forskellige transportmekanismer ansvarlige for denne kontrol er blevet<br />

undersøgt i de fleste dyregrupper, men vores viden om hvorledes dyr håndterer<br />

osmotisk stress, er stadig ufuldstændig. I dette arbejde er osmoregulatoriske<br />

fænomener blevet undersøgt i såvel vertebrater som invertebrater, omend der<br />

hovedsageligt er blevet fokuseret på dynamikkerne af vand og opløste stoffer i<br />

Tardigrada. Eksempelvis blev sammensætningen af uorganiske ioner undersøgt<br />

i flere forskellige arter af bjørnedyr, hvilket afslørede, at de relative bidrag af<br />

uorganiske ioner til den totale osmotiske koncentration overordnet set er ens i<br />

bjørnedyr og nært beslægtede dyregrupper. Desuden blev det udledt, at<br />

kryptobiotiske bjørnedyr (arter, der er i stand til indtræde i et stadie af latent<br />

liv) indeholder en stor andel af organiske osmolytter. Mekanismerne for<br />

aniontransport blev undersøgt farmakologisk i en marin art af bjørnedyr, og<br />

sammenlignet med de tilsvarende mekanismer i insekter. I bjørnedyret blev den<br />

organiske aniontransport lokaliseret til epitelet i midttarmen, og viste sig at<br />

være en aktiv transport, med en farmakologisk profil svarende til den i insekter.<br />

Bjørnedyr kan overleve et bredt spektrum af osmotiske miljøer (semiterrestrisk,<br />

limnisk og marine habitater), hvorfor evnen til at volumen- og osmoregulere<br />

blev undersøgt. Disse studier demonstrerede at bjørnedyr kan regulere den<br />

samlede kropsvolumen under såvel hypo- som hyperosmotiske forhold, samt<br />

indikerede at hyperregulering kan være en gennemgående tendens hos<br />

eutardigrader. Indeværende arbejder har altså bidraget til at fastslå bjørnedyr<br />

som en vigtig eksperimentel gruppe, hvori centrale fysiologiske spørgsmål kan<br />

besvares, heriblandt aspekter af ionregulering og osmoregulering.<br />

29


Acknowledgements<br />

During the course of my PhD studies I was fortunate to have the generous help<br />

of many great people, which has made this work possible.<br />

First and foremost, I thank my supervisor Nadja Møbjerg for being an<br />

unwavering source of support, encouragement and guidance during past years,<br />

but also for constantly challenging me to develop intellectually as well as a<br />

researcher. I truly appreciate all you have done for me, and hope to be able to<br />

continue our collaboration in the future.<br />

I also greatly appreciate the support of Reinhardt Møbjerg Kristensen<br />

(close to being my second supervisor) who initially introduced me to strange<br />

world of tardigrades, and who took me under his wing, offering me guidance<br />

and support (and equally important coffee). You have had enormous impact on<br />

my education as a young researcher, and I have sincerely enjoyed working with<br />

you.<br />

I have had the great fortune of working with many wonderful<br />

colleagues at University of Copenhagen, both at The August Krogh Center and<br />

the Natural History Museum who all are warmly thanked. Thanks to you it has<br />

been a pleasure coming to work every day. I am especially grateful to Erik<br />

Hviid Larsen with whom I shared many interesting discussions, and who was<br />

an endless source of inspiration. Aslak Jørgensen is thanked for great<br />

collaborations and discussions over the years as well as for critical review of<br />

my thesis. Dennis Krogh Persson is thanked for being a good colleague, but<br />

more importantly, a good friend. Since we started this journey in research<br />

together, I have come to depend on our collaborations and camaraderie, which<br />

has been an invaluable help along the way (by the way, it is your turn to buy<br />

cake). I thank Jette Lyby Michelsen for technical assistance over the years.<br />

During my visits to Roskilde University I was fortunate to collaborate<br />

with several great people. I especially thank Hans Ramløv for allowing me to<br />

come work in his laboratory (on several occasions), which scientifically turned<br />

out to be quite fruitful. I have particularly enjoyed our conversations and<br />

discussions, and I thank you for both your encouragement and support. I also<br />

thank Kristine Wulff Larsen for unparalleled project teamwork as well as our<br />

many conversations on everything and anything. It was an absolute pleasure,<br />

30


which I hope to repeat some time in the future. Peter Westh is thanked for<br />

allowing me to work in his laboratory and for his enormous expertise and help.<br />

Perhaps most importantly I thank the help and support of family and<br />

friends. I especially thank my parents and parent in-laws for their invaluable<br />

support during difficult times, and for helping me and my family in more ways<br />

than I care to mention. You’re the best! Lastly I thank my girlfriend Iben Rønn<br />

Veland for helping me, supporting me, tolerating me and loving me, but mostly<br />

for taking care of our son when I was working – none of this was possible<br />

without your help! I also thank my little son Hannibal for tolerating that I had<br />

to work sometimes, and couldn’t be home when you wanted me to. I am<br />

looking forward to making up for lost time. I love you both more than you<br />

know. Thank you.<br />

This work was funded by the 2008 Faculty of Science, University of<br />

Copenhagen Freja-Programme.<br />

31


References<br />

Anderson, E. (1960). The ultramicroscopic structure of the reptilian kidney. J Morph 106:205-<br />

240.<br />

Aguinaldo, A.M.A., Turbeville, J.M., Linford, L.S., Rivera, M.C., Garey, J.R., Raff, R.A.<br />

& Lake, J.A. (1997). Evidence for a clade of nematodes, arthropods and other moulting<br />

animals. Nature 387 (6632):489-493.<br />

Beyenbach, K. W. & Piermarini P. M. (2011). Transcellular and paracellular pathways of<br />

transepithelial fluid secretion in Malpighian (renal) tubules of the yellow fever mosquito<br />

Aedes aegypti. Acta Physiol 202:387-407.<br />

Bertolani, R (2001). Evolution of the reproductive mechanisms in tardigrades – a review. Zool<br />

Anz 240:247-252.<br />

Campbell, L.I., Rota-stabelli, O., Edgecombe, G.D., Marchioro, T., Longhorn, S.J.,<br />

Telford, M.J., Philippe, H., Rebecchi, L., Peterson, K.J. & Pisani, D. (2011).<br />

MicroRNAs and phylogenomics resolve the relationships of Tardigrada and suggest that<br />

velvet worms are the sister group of Arthropoda. PNAS 108:15920-15924.<br />

Campiglia, S. (1976). The blood of Peripatus acacioi Marcus & Marcus (Onychophora) – III.<br />

The ionic composition of the hemolymph. Comp Biochem Physiol 54(A):129-133.<br />

Coast, G. (2007). The endocrine control of salt balance in insects. Gen Comp Endocrinol<br />

152:332-338.<br />

Crowe, J. H. & Madin, K. A. (1974). Anhydrobiosis in tardigrades and nematodes. Trans Am<br />

Microsc Soc 93:513-524.<br />

Dantzler, W. H. (2002). Renal organic anion transport: a comparative and cellular perspective.<br />

Biochim Biophys Acta 1566:169-181.<br />

Dawson, D. C. & Liu, Xuehong (2009). Osmoregulation: some principles of water and solute<br />

transport. In: David H. Evans (ed) Osmotic and ionic regulation: cells and animals. CRS<br />

Press, Florida, USA.<br />

Degma, P., Bertolani, R. & Guidetti, R. 2009-2012. Actual checklist of Tardigrada species.<br />

Ver. 20: 17-01-2012. http://www.tardigrada.modena.unimo.it/miscellanea/Actual checklist<br />

of Tardigrada.pdf<br />

Degma, P. & Guidetti, R. (2007). Notes to the current checklist of Tardigrada. Zootaxa,<br />

1579:41-53.<br />

Dewel, R. A. & Dewel, W. C. (1979). Studies on the tardigrades. J Morphol 161:79-110.<br />

Diehl, W. J. (1986). Osmoregulation in echinoderms. Comp Biochem Physiol A 84:199-205.<br />

Dow, J. A. T. & Davies, S. A. (2006). The Malpighian tubule: Rapid insights from postgenomic<br />

biology. J Insect Physiol 52:365-378.<br />

Dowd, W. W., Harris, B. N. Chech Jr, J. J. & Kültz, D (2010). Proteomic and physiological<br />

responses of leopard sharks (Triakis semifasciata) to salinity change. J Exp Biol 213:210-<br />

224.<br />

32


Dunn, C.W., Hejnol, A., Matus, D.Q., Pang, K., Browne, W.E., Smith, S.A., Seaver, E.,<br />

Rouse, G.W. Obst, M., Edgecombe, G.D., Sørensen, M.V., Haddock, S.H.D.,<br />

Schmidt-Rhaesa, A., Okusu, A., Kristensen, R.M. & Wheeler, W.C. (2008). Broad<br />

phylogenomic sampling improves resolution of the animal tree of life. Nature 452:745-<br />

750.<br />

Edgecombe, G. D. (2010). Arthropod phylogeny: An overview from the perspective of<br />

morphology, molecular data and the fossil record. Arthropod Struct Dev 39:74-87.<br />

Essig, A. (1968). The “pump-leak“ model and exchange diffusion. Biophys J 8(1):53-63.<br />

Gabriel, W. N. & Goldstein, B. (2007). Segmental expression of Pax3/7 and Engrailed<br />

homologs in tardigrade development. Dev Genes Evol 217:421-433.<br />

George, R. L., Wu, X., Huang, W., Fei, Y. J., Leibach, F. H. & Ganapathy, V. (1999).<br />

Molecular cloning and functional characterization of a polyspecific organic anion<br />

transporter from Caenorhabditis elegans. J Pharmacol Exp Ther 291:596-603.<br />

Gilles, R. & Delpire, E. (1997). Variations in salinity, osmolarity and water availability. In:<br />

Dantzler, W. H. (ed.). Handbook of comparative physiology. New York: Oxford<br />

University Press, USA.<br />

Glasheen, J. S. & Hand, S. C. (1988). Anhydrobiosis in embryos of the brine shrimp Artemia:<br />

characterization of metabolic arrest during reductions in cell associated water. J Exp Biol<br />

135:363-380.<br />

Goldstein, B. & Blaxter, M. (2002). Tardigrades. Curr Biol 12:R475.<br />

Guidetti, R. & Bertolani, R. (2005). Tardigrade taxonomy: an updated check list of the taxa<br />

and a list of characters for their identification. Zootaxa 845:1-46.<br />

Guidetti, R., Bertolani, R. & Nelson, D. R. (1999). Ecological and faunistic studies on<br />

tardigrades in leaf litter of beech forest. Zool Anz 238:215-223.<br />

Hallas T, Yeates G (1972) Tardigrades of the soil and litter of a Danish beech forest.<br />

Pedobiologia 12:287-304.<br />

Hincha, D. K. & Hagemann, M. (2004). Stabilization of model membranes during drying by<br />

compatible solutes involved in the stress tolerance of plants and microorganisms. Biochem<br />

J 383:277-283.<br />

Hobsen, A. D., Stephenson, W. & Eden A. (1952). Studies on the physiology of Ascaris<br />

lumbricoides: II. The inorganic composition of the body fluid in relation to that of the<br />

environment. J Exp Biol 29:22-29.<br />

Jönson, K. I. (2003). Population density and species composition of moss-living tardigrades in<br />

a boreo-nemoral forest. Ecography 26:356-364.<br />

Jørgensen, A., Faurby, S., Hansen, J. G., Møbjerg, N. & Kristensen, R. M. (2010).<br />

Molecular phylogeny of Arthrotardigrada (Tardigrada). Mol Phylogenet Evol 54:1006-<br />

1015.<br />

Jørgensen, A., Møbjerg, N. & Kristensen, R. M. (2007). A molecular study of the tardigrade<br />

Echiniscus testudo (Echiniscidae) reveals low DNA sequence diversity over a large<br />

geographical area. Proceedings of the Tenth Internationl Symposium on Tardigrada. J<br />

limnol 66:77-83.<br />

Liggins, G. W. & Grigg, G. C. (1985). Osmoregulation of the cane toad Bufo marinus, in salt<br />

water. Comp Biochem Physiol A 82:613-619.<br />

Macallum, A. B. (1910). The inorganic composition of the blood in vertebrates and<br />

invertebrates, and its origin. Proc R. Soc Lond B 82(559):602-624.<br />

33


Maddrell, S. H. P. & Phillips, J. E. (1975). Secretion of hypo-osmotic fluid by the lower<br />

Malpighian tubules of Rhodnius prolixus. J Exp Biol 62:671-683.<br />

Martinez, E. A. (1975). Marine meiofauna of a New York City beach, with particular<br />

reference to Tardigrada. Est Coast Mar Sci 3:337-348.<br />

Morgan, C. I. (1977). Population dynamics of two species of Tardigrada, Macrobiotus<br />

hufelandii (Schultze) and Echiniscus (Echiniscus) testudo (Doyère), in roof moss from<br />

Swansea. J Anim Ecol 46:263-279.<br />

Møbjerg, N & Dahl, C. (1996). Studies on the morphology and ultrastructure of the<br />

Malpighian tubules of Halobiotus crispae Kristensen, 1982 (Eutardigrada). Zool J Linn<br />

Soc 116:85-99.<br />

Møbjerg, N., Jespersen, Å & Wilkinson, M. (2004). Morphology of the kidney in the West<br />

African caecilian, Geotrypetes seraphini (Amphibia, Gymnophiona, Caeciliidae). J Morph<br />

262(2):583-607.<br />

Nelson, D. R. (2002). Current status of Tardigrada: evolution and ecology. Integr Comp Biol<br />

42(3):652-659.<br />

Normant, M., Kubicka, M., Lapucki, T., Czarnowski, W. & Michalowska M. (2005).<br />

Osmotic and ionic haemolymph concentration in the Baltic Sea amphipod Gammarus<br />

oceanicus in relation to water salinity. Comp Biochem Physiol A 141:94-99.<br />

O’Donnell, M. J. & Maddrell, S. H. P. (1995). Fluid reabsorption and ion transport by the<br />

lower Malpighian tubules of adult female Drosophila. J Exp Biol 198:1647-1653.<br />

Peaker, M (1971). Avian salt glands. Philos Trans R Soc Lond B 262:289-300.<br />

Pelzer, B., Dastych, H. & Greven, H. (2007). The osmoregulatory/excretory organs of the<br />

glacier-dwelling eutardigrade Hypsibius klebelsbergi Mihelcic, 1959 (Tardigrada). Mitt.<br />

hamb. Zool Mus Inst 104:61-72.<br />

Ramazzotti, G. & Maucci W. (1983). Il Phylum Tardigrada. Terza edizione riveduta e<br />

corretta. Mem Insti Ital Idro Dott Marco De Marchi 41:1-1012.<br />

Reilly, D., Cramp, R. L., Wilson, J. M., Campbell, H. A. & Franklin, C. E. (2011).<br />

Branchial osmoregulation in the euryhaline bull shark, Charcharhinus leucas: a molecular<br />

analysis of ion transporters. J Exp Biol 214:2883-2895.<br />

Renaud-Mornant, J. (1982). Species diversity in marine Tardigrada. In: Nelson (Ed.)<br />

Proceedings of the 3rd international symposium on Tardigrada. East Tennessee State<br />

University Press, pp. 149-178.<br />

Reynolds, S. E. & Bellward, K. (1989). Water balance in Manduca sexta caterpillars: water<br />

recycling from the rectum. J Exp Bio 141:33-45.<br />

Ricci, C., Melone, G. Santo, N. & Caprioli, M. (2003). Morphological response of a Bdelloid<br />

Rotifer to dessication. J Morphol 257:246-253.<br />

Riegel, J. A. (1970). A new model of transepithelial fluid movement with detailed application<br />

to fluid movement in the crayfish antennal gland. Comp Biochem Physiol 36:403-410.<br />

Robertson, J. D. (1949). Ionic regulation in some marine invertebrates. J Exp Biol 26:182-<br />

200.<br />

Rota-Stabelli O, Kayal E, Gleeson D, Daub J, Boore J, Telford M, Pisani D, Blaxter M,<br />

Lavrov D. (2010). Ecdysozoan mitogenomics: evidence for a common origin of the<br />

legged invertebrates, the Panarthropoda. Genome Biol Evol 2:425-440.<br />

Sardella, B. A., Baker, D. W. & Brauner, C. J. (2009). The effects of variable water salinity<br />

and ionic composition on the plasma status of the Pacific Hagfish (Eptatretus stoutii). J<br />

Comp Physiol B 179:721-728.<br />

34


Sarver, R. G., Flynn, M. A. & Holliday, C. W. (1994). Renal Na, K-ATPase and<br />

osmoregulation in the crayfish Procambarus clarkii. Comp Biochem Physiol (A) 107:349-<br />

356.<br />

Schluter, D. (2009). Evidence for ecological speciation and its alternative. Science 323:737-<br />

741.<br />

Schmidt-Nielsen, K. (1963). Osmotic regulation in higher vertebrates. Harvey lect 58:53-93.<br />

Schmidt-Rhaesa A, Kulessa J. (2007). Muscular architecture of Milnesium tardigradum and<br />

Hypsibius sp. (Eutardigrada, Tardigrada) with some data on Ramazzottius oberhaeuseri.<br />

Zoomorphol 126:265-281.<br />

Sutcliffe, D. W. (1962). The composition of haemolymph in aquatic insects. J Exp Biol<br />

39:325-343.<br />

Suzuki, A. C. (2003). Life history of Milnesium tardigradum Doyère (Tardigrada) under a<br />

rearing environment. Zool Sci 20:49-57.<br />

Torrie, L. S., Radford, J. C., Southall, T. D., Kean, L., Dinsmore, A. J., Davies, S. A. &<br />

Dow, J. A. T. (2004). Resolution of the insect ouabain paradox. PNAS 101: 13689-13693.<br />

Uhía, E. & Briones, M. J. I. (2002) Population dynamics and vertical distribution of<br />

enchytraeids and tardigrades in response to deforestation. Acta Oecolog 23:349-359.<br />

Van Weel, P. B. (1957). Observations on the osmoregulation in Aplysia juliana Pease<br />

(Aplysiidae, Mollusca). Z vergl Physiol 39:492-506.<br />

Walz, B. (1974). The Fine Structure of Somatic Muscles of Tardigrada. Cell Tiss Res 149:81-<br />

89.<br />

Weglarska, B. (1987). Studies on the excretory system of Isohypsibius granulifer Thulin<br />

(Eutardigrada). In: Biology of tardigrades. Selected symposia and monographs (1) (ed.<br />

Bertolani, R.), pp. 15-24. Modena: U.Z.I. Mucchi.<br />

Whittamore, J. M. (2012). Osmoregulation and epithelial water transport: lessons from the<br />

intestine of marine teleost fish. J Comp Physiol B 182:1-39.<br />

Wilder, M. N., Ikuta, K., Atmomarsono M., Hatta, T. & Komuro, K. (1998). Changes in<br />

osmotic and ionic concentrations in the hemolymph of Macrobrachium rosenbergii<br />

exposed to varying salinities and correlation to ionic and crystalline composition of the<br />

cuticle. Comp Biochem Physiol A 119:941-950.<br />

Williams, J. C. & Beyenbach, K. W. (1983). Differential effects of secretagogues on Na and<br />

K secretion in the Malpighian tubules of Aedes aegypti. J Comp Physiol B 149:511-517.<br />

Yancey, P. H. (2005). Organic osmolytes as compatible, metabolic and counteracting<br />

cytoprotectants in high osmolarity and other stresses. J Exp Biol 208:2819-2830.<br />

Zantke, J., Wolff, C. & Scholtz, G. (2008). Three-dimensional reconstruction of the central<br />

nervous system of Macrobiotus hufelandi (Eutardigrada, Parachela): implications for the<br />

phylogenetic position of Tardigrada. Zoomorphol 127:21–36.<br />

Zhao, H. (2005). Effect of ions and other compatible solutes on enzyme activity, and its<br />

implication for biocatalysis using ionic liquids. J Mol Catal B: Enzymat 37:16-25.<br />

35


36<br />

UNIVERSITAS<br />

HAFNIENSIS<br />

2012


Paper I


Inorganic ion composition in Tardigrada: cryptobionts contain large<br />

fraction of unidentified organic solutes<br />

<strong>Kenneth</strong> Agerlin <strong>Halberg</strong> 1,* , Kristine Wulff Larsen 2 , Aslak Jørgensen 3 , Hans Ramløv 2 and<br />

Nadja Møbjerg 1<br />

1 Department of Biology, the August Krogh Centre, University of Copenhagen, Universitetsparken 13, DK-<br />

2100 Copenhagen Ø, Denmark, 2 Department of Nature, Systems and Models, University of Roskilde,<br />

Universitetsvej 1, DK-4000 Roskilde, Denmark and 3 Laboratory of Molecular Systematics, Natural History<br />

Museum of Denmark, University of Copenhagen, Sølvgade 83, DK-1307 Copenhagen K, Denmark<br />

*Author for correspondence (kahalberg@bio.ku.dk)<br />

Submitted June 2012<br />

SUMMARY<br />

Tardigrades are a group of micrometazoans known to tolerate extreme environmental stress. Significant<br />

efforts have been devoted to the field, however; mechanisms explaining the extreme adaptations found<br />

among tardigrades is still lacking. Here we present data on the inorganic ion composition and total<br />

osmotic concentration of five different species of tardigrades (E. testudo, M. tardigradum, R. coronifer, M.<br />

cf. hufelandi and H. crispae) using high-performance anion-exchange chromatography and nanoliter<br />

osmometry. Quantification of the ionic content indicates that Na + and Cl - are the principle inorganic ions<br />

in tardigrade fluids, albeit substantial concentrations of K + , NH + 4 , Ca 2+ , Mg 2+ , F - 2-<br />

, SO 4 and PO 3- 4 also<br />

were detected. In limno-terrestrial tardigrades, the respective ions are concentrated by a large factor<br />

compared to that of the external medium (Na + , ×70-800; K + , ×20-90; Ca 2+ and Mg 2+ , ×30-200; Cl - , ×20-50;<br />

SO 2- 4 , ×30-150), whereas in the marine species H. crispae Na + , Cl - 2-<br />

and SO 4 are almost in ionic<br />

equilibrium with (brackish) salt water, while K + , Ca 2+ and Mg 2+ are slightly concentrated (×2-10).<br />

However, there is an anion deficit of ~120 mEq/l in M. tardigradum and H. crispae, indicating that there<br />

are ionic components that remain unidentified in these species. Body fluid osmolality ranged from 361±49<br />

in R. coronifer to 961±43 mOsm/kg in H. crispae. Concentrations of most inorganic ions are largely<br />

identical between active and dehydrated groups of R. coronifer, suggesting that this tardigrade does not<br />

exclude large amounts of ions during dehydration. The large osmotic and ionic gradients maintained by<br />

both limno-terrestrial and marine species are indicative of a powerful ion-retentive mechanism in<br />

Tardigrada. Moreover, our data indicates that cryptobiotic tardigrades contain a large fraction of<br />

unidentified organic osmolytes, the identification of which is expected to provide increased insight into<br />

the phenomenon of cryptobiosis.<br />

Key words: tardigrades, inorganic ions, ion chromatography, nanoliter osmometry, organic osmolytes,<br />

cryptobiosis<br />

INTRODUCTION<br />

Tardigrades are a group of minute multi-cellular<br />

animals that are known to tolerate extreme<br />

environmental stress (Guidetti et al., 2010;<br />

Møbjerg et al., 2011). This capacity derives<br />

mainly from their ability to enter a state latent of<br />

life, i.e. cryptobiosis, in which their resistances to<br />

adverse environmental conditions are greatly<br />

increased (Møbjerg et al., 2011). In recent years,<br />

intensive research efforts have been devoted to the<br />

field, as the translational output associated with a<br />

detailed understanding of their complex stress<br />

biology is expected to include new methods for<br />

preserving and stabilizing biological materials<br />

(Wełnicz et al., 2010). Several advances have been<br />

made in our understanding of tardigrade stress<br />

responses, especially regarding i) the role of<br />

selective carbohydrates (trehalose), ii) the


K. A. <strong>Halberg</strong> and others<br />

differential expression of stress proteins (heat<br />

shock proteins and late embryogenesis abundant<br />

proteins), and iii) the identification of antioxidant<br />

defenses and DNA repair mechanisms<br />

(see reviews by Guidetti et al., 2010; Wełnicz et<br />

al., 2010; Møbjerg et al., 2011). However, a<br />

general mechanism explaining the extreme<br />

adaptations found among tardigrades is still<br />

lacking. Consequently, new approaches that may<br />

provide increased insight into the superior stress<br />

adaptations of tardigrades are greatly needed.<br />

Knowledge of the composition as well as<br />

concentrations of dissolved particles in internal<br />

fluids is fundamental to the understanding of<br />

basic physiological processes, such as fluid and<br />

electrolyte homeostasis, signal transduction and<br />

solute transport. Accordingly, such data have<br />

been provided for most major groups of animals<br />

(cnidarians, echinoderms, annelids, molluscs,<br />

crustaceans, insects, chelicerates, tunicates, fish,<br />

amphibians and mammals) more than half a<br />

century ago (Macallum, 1910; Robertson, 1949,<br />

1954; Sutcliffe, 1962; Hronoski & Armstrong,<br />

1977). However, practically nothing is known<br />

about the chemical composition of tardigrades,<br />

which has been a major obstacle to the<br />

understanding of the fluid and solute dynamics in<br />

these animals (<strong>Halberg</strong> et al., 2009b; Møbjerg et<br />

al., 2011; <strong>Halberg</strong> & Møbjerg, 2012). Questions<br />

relating to this area of research are especially<br />

important to address, if we wish to unravel the<br />

biological mechanisms mediating the unique<br />

tolerance to extreme desiccation (anhydrobiosis)<br />

– the most widespread form of cryptobiosis in<br />

Tardigrada.<br />

In the present study, we use a combination of<br />

high-performance anion-exchange chromatography<br />

(HPAEC) and nanoliter osmometry, to<br />

identify and quantify inorganic cations and anions<br />

present in tardigrade homogenates, and to<br />

measure the total osmotic concentrations of five<br />

different species of tardigrades, covering both a<br />

broad phylogenetic and habitat spectrum. Our<br />

study indicates that tardigrades possess powerful<br />

ion-retentive and osmoregulatory capacities, and<br />

that (only) cryptobiotic species contain a large<br />

fraction of organic solutes.<br />

MATERIALS AND METHODS<br />

Tardigrade sampling<br />

Specimens of Richtersius coronifer Richters,<br />

1908, Macrobiotus cf. hufelandi C.A.S. Schultze,<br />

1834 and Milnesium tardigradum Doyère, 1840<br />

were extracted from moss collected at Öland,<br />

Sweden, while Echiniscus testudo Doyère, 1840<br />

was found in moss collected at Nivå, Denmark.<br />

These species were extracted by washing the<br />

respective moss samples with tap water through<br />

six different sieves of progressively smaller mesh<br />

size, so as to concentrate the tardigrades and to<br />

remove large debris. Specimens of Halobiotus<br />

crispae Kristensen, 1982 were isolated from<br />

marine algae and sediment collected from<br />

Vellerup Vig, Denmark according to the method<br />

of <strong>Halberg</strong> and Møbjerg (2012). The total number<br />

of animals used in all experiments was 2220 R.<br />

coronifer, 326 M. tardigradum, 630 M. cf.<br />

hufelandi, 426 E. testudo and 268 H. crispae.<br />

Inorganic cation and anion analysis<br />

The dominant cations and anions present in the<br />

different tardigrade species were determined by<br />

HPAEC using a Metrohm chromatography system<br />

(830 IC interface, 818 IC pump, 819 IC<br />

conductivity detector, columns C4 - 150/4.0<br />

(cations) and A supp 5 150/4.0 (anions); Metrohm,<br />

Herisau, Switzerland). The eluents (mobile<br />

phases) were made according to manufacturer’<br />

instructions: For cations, the eluent consisted of<br />

0.7 mM C 7 H 5 NO 4 + 1.7 mM (65%) HNO 3 . For<br />

anions, it consisted of 3.2 mM Na 2 CO 3 + 1.0 mM<br />

NaHCO 3 . The eluents were filtered (mesh size: 45<br />

µm) prior to use. The analysis settings employed<br />

were a flow-rate of 0.9 ml/min (cations) and 0.7<br />

ml/min (anions) with a pressure of ~6.4 MPa.<br />

Cation analyses were performed non-suppressed,<br />

whereas anion detection was conducted using<br />

chemical suppression. Fluka multi-element cation<br />

and anion standards (Sigma-Aldrich, St. Louis,<br />

MO, USA) were used to construct calibration<br />

curves for the respective ions bracketing the<br />

concentration range of interest. Based on these<br />

calibration curves, the ion chromatography (IC)<br />

software (IC Net 2.3, Metrohm, Herisau,<br />

Switzerland) calculated the ion concentrations of<br />

all subsequent samples (mg/l), which were<br />

recalculated to a different unit of concentration<br />

(mM) and adjusted according to the appropriate<br />

dilution factor (see below). Representative<br />

chromatograms of both the cationic and anionic<br />

fractions are shown for all investigated species<br />

(Figs. 1, 2). The empirically determined elution<br />

order and retention times of the investigated ions<br />

were Na + (t R = 5.37 min), NH 4 + (t R = 6.03 min),


Inorganic ion composition in Tardigrada<br />

Fig. 1. Representational chromatograms revealing the principal inorganic cations present in each investigated species:<br />

1, sodium; 2, ammonium; 3, potassium; 4, calcium; 5, magnesium; nOAp, negative organic acid peak. Stitched square<br />

indicates an unidentified compound (t R = 10.36 min) that increases app. two-fold in absolute concentration in<br />

dehydrated animals of Richtersius coronifer (data not shown). Column, Metrohm C4-150/4.0; mobile phase, 0.7 mM<br />

dipicolinic acid + 1.7 mM (65%) nitric acid; flow-rate, 0.9 ml/min; Conductivity detector without suppression.<br />

Injection volume, 60 µl.<br />

K + (t R = 7.73 min), Ca 2+ (t R = 18.12 min), Mg 2+<br />

(t R = 23.17 min), F - (t R = 4.05 min), Cl - (t R = 6.01<br />

min), PO 4 3- (t R = 14.74 min) and SO 4 2- (t R = 16.58<br />

min).<br />

Sample preparation<br />

Following extraction, specimens were washed<br />

repeatedly with ddH 2 O (Halobiotus crispae was<br />

washed in filtered salt water; SW, 20 ‰), and<br />

subsequently transferred, using an Irwin loop, to<br />

sample tubes containing cation eluent (75-100<br />

µl); samples dissolved in cation eluent allows for<br />

a more precise quantification of cations due to<br />

increased signal to noise ratio (pers. comm.;<br />

Metrohm Nordic, Denmark). Prior to transfer,<br />

surface water was removed by blotting the<br />

animals with tissue paper in an attempt to avoid<br />

unwanted dilution of the samples. Pilot<br />

experiments revealed that this process was critical<br />

to acquire reproducible data, and was accordingly<br />

performed as fast and uniform as possible. A total<br />

of 40-225 animals were transferred to each test<br />

tube and the sample was subsequently<br />

homogenized using a sterile plastic pestle; great<br />

care was taken to ensure complete homogenization<br />

(visually confirmed at 50× magnification), and the<br />

pestle was subsequently rinsed with a small<br />

volume of cation eluent to ensure total transfer of<br />

ions to the test tube. The number of animals per<br />

sample (N) varied according to species size and<br />

availability (Tables 1, 2). The entire sample was<br />

then centrifuged (10 min at 5600 rpm) to remove<br />

solid particles (e.g. cuticle fragments), and the<br />

supernatant was filtered (mesh size: 0.20 µm)<br />

using a single use syringe filter (Sartorius AG,<br />

Göttingen, Germany). The samples were<br />

subsequently frozen at -20 ºC, if not quantified<br />

immediately. A total of five to seven samples were<br />

prepared for each species (Table 2).


K. A. <strong>Halberg</strong> and others<br />

Fig. 2. Representational chromatograms revealing the principal inorganic anions present in each investigated species:<br />

6, fluoride; 7, chloride; 8, phosphate; 9, sulfate; W, negative water peak. Stitched square indicates an unidentified<br />

compound (t R = 7.24 min) that increases app. two-fold in absolute concentration in dehydrated animals of Richtersius<br />

coronifer (data not shown). Column, Metrohm A supp 5 150/4.0; mobile phase, 3.2 mM sodium carbonate + 1.0 mM<br />

sodium hydrogen carbonate; flow-rate, 0.7 ml/min; Conductivity detector with chemical suppression. Injection<br />

volume, 60 µl.<br />

The ionic concentration and composition of<br />

the external media from the different habitats i.e.<br />

moss water (MW) and SW, were additionally<br />

determined. Moss samples were rehydrated in<br />

ddH 2 O for several hours, and MW samples were<br />

subsequently collected from between the leafcovered<br />

stems. SW samples were prepared by<br />

diluting SW (1:200) collected at the locality.<br />

Samples were quantified in triplets using both<br />

vapor pressure osmometry (Vapro 5520, Wescor<br />

inc., UT, USA) and HPAEC (Table 3).<br />

In order to document whether changes in<br />

inorganic ion content occur during dehydration<br />

from an active to a cryptobiotic state, samples of<br />

dehydrated Richtersius coronifer were<br />

additionally prepared. Groups of 75 animals were<br />

transferred to each sample tube, and excess water<br />

was removed by blotting the animals with tissue<br />

paper. The animals were subsequently allowed to<br />

dehydrate over the ensuing 24 h at ambient<br />

temperature and humidity. Following complete<br />

dehydration, tissue paper saturated with ddH 2 O<br />

was used to rinse the surface of the animals, and a<br />

dry tissue paper was used to remove excess<br />

moisture. This was done in order to remove<br />

potential solutes extruded on the surface of the<br />

animals – the surface of some animals was<br />

inaccessible due to animal clumping, and therefore<br />

could not be rinsed. A volume of 90 µl of cation<br />

eluent was added, and the animals were<br />

immediately homogenized. The samples were then<br />

prepared as described above with six samples<br />

prepared in total (Table 2). Data (mg/l) from this<br />

experiment was directly compared to that of<br />

hydrated animals (Table 4; Fig. 3B), as both sets<br />

of samples contained near identical number of<br />

animals per unit volume (i.e. 0.81 and 0.83<br />

animals/µl eluent respectively), which<br />

circumvented the need for recalculations (see<br />

below). In order to test whether R. coronifer<br />

actually produced viable tuns during the<br />

abovementioned conditions, post-cryptobiotic


Inorganic ion composition in Tardigrada<br />

survival was assessed. Using four groups of 50<br />

specimens, a survival rate of 94 ± 4% (mean ±<br />

s.d.) was observed, which is comparable to the<br />

maximally reported survival rate of R. coronifer<br />

dehydrated on Whatman filters (Persson et al.,<br />

2010).<br />

Calculation of ion concentrations<br />

The IC software expressed the integration of<br />

peaks as a concentration (mg/l), which was<br />

recalculated using the respective molecular<br />

weights of each compound to a different<br />

concentration (mM) prior to adjusting for the<br />

dilution factors.<br />

The volume of each investigated species was<br />

calculated according to an adjusted method of<br />

<strong>Halberg</strong> et al. (2009b). In brief, micrographs were<br />

taken of N = 20 animals of each species, using a<br />

digital camera (C-5050, Olympus, Japan)<br />

mounted on an Olympus BX 51microscope<br />

(Olympus, Japan), and median length (h) and<br />

width (2r) of the trunk and legs were measured.<br />

Approximating the geometric shape of the trunk<br />

and legs as a cylinder, and adjusting the volume<br />

of liquid according to the gravimetrically<br />

measured water content (based on Westh and<br />

Kristensen, 1992; <strong>Halberg</strong> et al., 2009b), the fluid<br />

volume of an individual tardigrade of each<br />

species was calculated using equation (1).<br />

(Eq 1) V individual = π(r 2 trunkh trunk + 8r 2 legh leg ) × W<br />

Where V individual is the volume of an individual<br />

tardigrade, r is the radius and h the length of the<br />

trunk and hind legs respectively, while W (0.72 i.e.<br />

mean fractional water content of R. coronifer and<br />

H. crispae) is the gravimetrically measured<br />

fractional water content. Using these data (Table<br />

1), the total tardigrade test volume was calculated<br />

by multiplying the volume of an individual with<br />

the number of animals included in the sample<br />

according to equation (2).<br />

(Eq 2) V total = V individual × N<br />

V total is the total tardigrade test volume, and N the<br />

number of animals included in the sample. Lastly,<br />

the concentrations of the dominant cations and<br />

anions in the investigated species of tardigrades<br />

were calculated by multiplying the measured ion<br />

concentrations with the dilution factor, which was<br />

calculated according to equation (3).<br />

(Eq 3) D = F / V total<br />

Where D is the dilution factor and F is the final<br />

volume (i.e. volume of cation eluent the<br />

tardigrades were transferred to + V total ). Sample<br />

information for the respective species is listed in<br />

Table 2.<br />

Nanoliter osmometry<br />

The total osmotic concentration of tardigrades<br />

from each investigated species was estimated<br />

using nanoliter osmometry. This was done in order<br />

to determine the fraction that the identified<br />

Table 1. Volume estimations. Mean values of length (h) and width (2r) of the trunk and legs (N = 20 animals), as well<br />

as the calculated volume (Eq 1), of each investigated species. W is the average of the gravimetrically determined<br />

water content (72%) of Richtersius coronifer (Westh and Kristensen, 1992) and Halobiotus crispae (<strong>Halberg</strong> et al.,<br />

2009). Data are expressed as mean ± s.d.


K. A. <strong>Halberg</strong> and others<br />

inorganic ions constitute of the total osmotic<br />

concentration in each species, and to provide an<br />

independent verification of our HPAEC data, i.e.<br />

total osmotic concentration should be higher than<br />

the accumulated concentration of the respective<br />

inorganic ions (Table 3). Using the same<br />

procedure for removing excess water as described<br />

above, individual specimens were transferred into<br />

sample oil wells (loading oil type B; cST=1250 ±<br />

10%; Cargille laboratories, Cedar grove, NJ<br />

07009, USA) of a calibrated nanoliter osmometer<br />

(Clifton Technical Physics, Hartford, NY, USA),<br />

and the osmolality (mOsm/kg) was determined by<br />

freezing point depression (FPD = 1.858<br />

°C/Osmol). Six to ten animals of each species<br />

were used in this experiment (Table 3).<br />

Statistics<br />

Significant changes in the individual inorganic<br />

ion concentrations between active and<br />

cryptobiotic animals of Richtersius coronifer<br />

were tested using an unpaired, two-sample t-test<br />

with significance levels of P>0.05 (not<br />

significant, NS), P


Inorganic ion composition in Tardigrada<br />

Figure 3. Graphical representation of the respective ionic contributions to total osmotic concentration. A.<br />

Concentrations (mM) of the respective cations and anions measured in each investigated species, as well as the<br />

corresponding total osmotic concentration (mOsm/kg), as measured by nanoliter osmometry (see also Table 3). The<br />

blank area represents the osmotic deficit (OD), i.e. other solutes. The phylogenetic position and habitat preference of<br />

each species is listed. The light micrographs of the animals are shown to scale (see Table 1 for average species size).<br />

B. Concentrations (mg/l) of the respective cations and anions measured in hydrated, active specimens compared to<br />

dehydrated cryptobiotic animals of Richtersius coronifer (see also Table 4). Data are expressed as mean ± s.d.


K. A. <strong>Halberg</strong> and others<br />

Halobiotus crispae containing the lowest and<br />

highest concentrations respectively (Fig. 3A;<br />

Table 3). There are notable differences in the Na + /<br />

Cl - ratio between the animals, i.e., the ratio is less<br />

than unity in the limno-terrestrial herbivores E.<br />

testudo (0.61), R. coronifer (0.60) and M. cf.<br />

hufelandi (0.33), higher than unity in the limnoterrestrial<br />

predator M. tardigradum (1.19) and<br />

close to unity in the marine herbivore H. crispae<br />

(0.92).<br />

Compared with Na + and Cl - , generally, the<br />

[K + ] is relatively low in all species, ranging from<br />

19-73 mM (Tab. 3). Thus, the Na + / K + ratio is<br />

higher than unity in E. testudo (2.39), M.<br />

tardigradum (2.05), R. coronifer (1.32) and H.<br />

crispae (6.12), however; lower than unity in M.<br />

cf. hufelandi (0.77). Interestingly, the relative<br />

contribution of K + to total osmotic concentration<br />

is low constituting


Inorganic ion composition in Tardigrada<br />

Table 3. Ionic composition and total osmotic concentration of the investigated species of tardigrades and the corresponding external media. Concentrations (mM) of<br />

cations and anions detected in each investigated species, as well as moss water (MS) and 20 ‰ salt water (SW) samples, in addition to the corresponding total osmotic<br />

concentration (mOsm/kg), as measured by nanoliter osmometry or vapor pressure osmometry, respectively. In addition, the osmotic deficits (calculated as the<br />

difference in ionic concentration, and total osmotic concentration), as well as the observed charge deficits (calculated as the difference between positive and negative<br />

charges) are listed; the polarity of the charge deficits is indicated in parenthesis. Numbers noted in brackets indicates the number of samples tested. Data are expressed<br />

as mean ± s.d.


K. A. <strong>Halberg</strong> and others<br />

In the present study we provide data on the ionic<br />

composition of five different species of<br />

tardigrades covering a large phylogenetic<br />

spectrum. Our study is represented by members<br />

of Heterotardigrada (Echiniscoidea) and<br />

Eutardigrada (Apochela and Parachela), four<br />

evolutionary distant families (Echiniscidae,<br />

Milnesiidae, Macrobiotidae and Hypsibiidae), as<br />

well as both limno-terrestrial and marine habitats.<br />

Accordingly, we will discuss the ionic<br />

compositions of the respective species in relation<br />

to systematic position and habitat preference, as<br />

well as make comments on our data in relation to<br />

hemolymph composition in representatives of<br />

phylogenetically related groups (i.e. Arthropoda<br />

and Onychophora).<br />

Echiniscus testudo (Heterotardigrada:<br />

Echiniscidae) belongs to another evolutionary<br />

lineage than the other tardigrades in the present<br />

study. Compared to limno-terrestrial members of<br />

Eutardigrada, the ionic composition of this<br />

heterotardigrade is characterized by a large<br />

contribution of Na + and Cl - (~45%), and a very<br />

low contribution of Mg 2+ 2-<br />

(0.4%), SO 4 (0.6%)<br />

and PO 3- 4 (1.4%), respectively (Tab. 5). The large<br />

contribution of Na + and Cl - to total osmotic<br />

concentration, which is comparable to that seen in<br />

the marine species H. crispae (Table 5), could<br />

reflect the supposed marine origin of tardigrades<br />

(Jørgensen et al., 2010). This hypothesis can be<br />

tested by data on members of the ‘ancient’ and<br />

exclusively marine Arthrotardigrada.<br />

The family Milnesiidae, represented by the<br />

predator Milnesium tardigradum, is currently<br />

considered the sister-group of all other<br />

eutardigrades (Guidetti et al., 2009). M.<br />

tardigradum contains the highest total osmotic- as<br />

well as ionic concentration among the limnoterrestrial<br />

species, with conspicuously high levels<br />

of both K + and Ca 2+ (Table 3). The high [K + ] in<br />

M. tardigradum compared to the phytophagous<br />

species is somewhat surprising, as e.g.<br />

carnivorous insects typically contain low levels of<br />

K + (Sutcliffe, 1962). Conversely, phytophagous<br />

tardigrades are known to feed on bryophytes high<br />

in K + and low in Na + (Smith, 1978), and were,<br />

analogous to phytophagous insects (Sutcliffe,<br />

1962), expected to reflect this relative ion<br />

composition in their extracellular body fluids.<br />

Interestingly, the relative ion contributions to<br />

total osmotic concentration suggests that M.<br />

tardigradum resembles the heterotardigrades more<br />

than the other eutardigrades (Table 5)<br />

Sutcliffe (1962, 1963) argued that definitive<br />

types of hemolymph are related to phylogenetic<br />

position within Insecta. Apart from small<br />

differences in absolute concentrations of Na + and<br />

PO 3- 4 , the ionic compositions and relative<br />

contributions of the different components in<br />

Richtersius coronifer and Macrobiotus cf.<br />

hufelandi (Eutardigrada: Macrobiotidae) are<br />

similar (Tables. 3 and 5). As a testable hypothesis,<br />

these similarities would suggest that the relative<br />

ion composition among the species relates to<br />

phylogeny and systematic position in Tardigrada.<br />

In contrast to the other groups of limno-terrestrial<br />

tardigrades, the inorganic content of R. coronifer<br />

and M. cf. hufelandi is characterized by a<br />

relatively small contribution of Na + (~2-7 times<br />

lower) and Cl - (~2-3 times lower), and conversely,<br />

2-<br />

a relatively large contribution of SO 4 (~4-10<br />

times higher). The physiological significance of<br />

these variations is unknown.<br />

Halobiotus crispae (Eutardigrada:<br />

Hypsibiidae) is a truly marine species, and is the<br />

species with the highest total concentration of both<br />

ions and total solutes measured. Na + and Cl -<br />

account for more than 50% of its total osmotic<br />

concentration. The divalent cations, Ca 2+ and<br />

Mg 2+ , are also detected in high concentrations,<br />

both absolute and relative. In contrast to the<br />

limno-terrestrial species, the total osmotic<br />

concentration of H. crispae is almost exclusively<br />

accounted for by the measured ionic<br />

concentrations, which becomes evident when<br />

considering the charge deficit indicated in Table 3.<br />

The contribution of the total diffusible ions to<br />

the total osmotic concentration of tardigrades is<br />

roughly similar to that of the hemolymph of<br />

arthropods, nematodes and onychophorans (Table<br />

5). In fact, as Na + predominantly is an<br />

extracellular ion, whereas K + and Ca 2+ mainly are<br />

intracellular ions, the ionic composition of<br />

tardigrade hemolymph is expected to resemble<br />

that of closely related animal groups, e.g. other<br />

members of Panarthropoda (see Table 5).<br />

It is relevant to compare concentrations of ions<br />

in tardigrade body fluids with those of the<br />

respective external media (Table 3). The strongest<br />

ability to concentrate ions is seen in limnoterrestrial<br />

species, which hyper-regulate by as<br />

much as ~350-750 mOsm/kg (Fig. 3A). The<br />

marine H. crispae maintains hemolymph osmotic


Inorganic ion composition in Tardigrada<br />

Table 4. Changes in ionic<br />

composition during dehydration.<br />

Concentrations (mg/l) of<br />

the respective cations and<br />

anions measured in hydrated,<br />

active specimens compared to<br />

dehydrated cryptobiotic animals<br />

of Richtersius coronifer. Both<br />

sets of samples contained near<br />

identical number of animals per<br />

unit volume (i.e. 0.8 animals/µl<br />

eluent), and were accordingly<br />

directly comparable. Numbers<br />

noted in brackets indicates the<br />

number of samples tested. Data<br />

are expressed as mean ± s.d.<br />

Significant difference in the<br />

concentration of the respective<br />

inorganic ion concentrations<br />

were tested using an unpaired,<br />

two-sample t-test with<br />

significance levels of P>0.05<br />

(not significant, NS), P


K. A. <strong>Halberg</strong> and others<br />

intracellular osmotic potential so that osmotic<br />

equilibrium between intra- and extracellular<br />

fluids is maintained. Moreover, organic<br />

osmolytes are known to stabilize macromolecular<br />

structures by direct interaction with proteins and<br />

membrane lipids (Crowe et al., 1987; Hincha and<br />

Hagemann, 2004; Yancey, 2005). As evidenced<br />

by the large osmotic deficits, which appears<br />

restricted to cryptobiotic tardigrades (Fig. 3A;<br />

Table 3), it is tempting to suggest that a large<br />

quantity of organic osmolytes are synthesized in<br />

these species, thus enabling the animals to<br />

respond quickly to decreases in external water<br />

potential. Such a strategy seems favorable in light<br />

of the continuous dehydration-rehydration cycles<br />

that may occur in limno-terrestrial habitats,<br />

additionally supported by the short time span<br />

(


Table 5. Total osmotic concentration (mOsm/kg) of the external medium and internal body fluids, as well as the osmotic contribution (%) of the respective ions to the<br />

internal concentration in each of the investigated species of tardigrades. Corresponding data on hemolymph concentration and composition of selected species of<br />

nematodes, crustaceans, insects and onychophorans is included for comparative purposes. TR, terrestrial; ─, not measured.<br />

Inorganic ion composition in Tardigrada


K. A. <strong>Halberg</strong> and others<br />

We would like to thank Anne Lise Maarup for technical<br />

assistance and Reinhardt Møbjerg Kristensen (Natural History<br />

Museum of Denmark) for loan of the Olympus BX 51<br />

microscope. Station Linné (Porten til Alvaret), Ölands Skogsby,<br />

Sweden is warmly thanked for accommodation during sampling<br />

of tardigrades. Funding came from the Carlsberg Foundation<br />

and the Freja-Programme (Faculty of Science, University of<br />

Copenhagen).<br />

FPD<br />

HPAEC<br />

IC<br />

MW<br />

SW<br />

LIST OF ABBREVIATIONS<br />

freezing point depression<br />

high-performance anion-exchange chromatography<br />

ion chromatography<br />

moss water<br />

salt water<br />

REFERENCES<br />

Campiglia, S. (1976). The blood of Peripatus acacioi<br />

Marcus & Marcus (Onychophora) – III. The ionic<br />

composition of the hemolymph. Comp. Biochem.<br />

Physiol., 54(A):129-133.<br />

Chen, T., Acker, J. P., Eroglu, A. flere authors<br />

(2001). Beneficial effect of intracellular trehalose on<br />

the membrane integrity of dried mammalian cells.<br />

Cryobiol. 43:168-181.<br />

Crowe, J. H., Crowe, L. M., Carpenter, J. F. and<br />

Wistrom, C. A. (1987). Stabilization of dry<br />

phospholipid bilayers and proteins by sugars.<br />

Biochem. J., 242:1-10.<br />

Bird, A. F., McClure, S. G. (1997). Composition of the<br />

stylets of the tardigrade, Macrobiotus cf.<br />

pseudohufelandi. Trans R Soc. S Aust., 121:43-50.<br />

Eagers, R. Y. (1969). Toxic properties of inorganic<br />

fluoride compounds. Elsevier, Amsterdam London<br />

New York, 382 pp.<br />

Granat, L. (1972). On the relation between pH and the<br />

chemical composition in atmospheric precipitation.<br />

Tellus XXIV, 6.<br />

Guidetti, R., Altiero, T. & Rebbecchi, L. (2010). On<br />

dormancy strategies in tardigrades. J. Insec.<br />

Physiol., 57(5):567-576.<br />

Guidette, R., Schill, R. O., Bertolani, R., Dandekar<br />

T. & Wolf, M. (2009). New molecular data for<br />

tardigrade phylogeny, with the erection of<br />

Paramacrobiotus gen. nov. J. Zool. Syst. Ecol. Res.,<br />

47(4):315-321.<br />

<strong>Halberg</strong>, K. A., Persson D., Møbjerg N., Wanninger<br />

A. & Kristensen R. M. (2009a). Myoanatomy of<br />

the Marine Tardigrade Halobiotus crispae<br />

(Eutardigrada: Hypsibiidae). J. Morphol., 270:996-<br />

1013.<br />

<strong>Halberg</strong>, K. A., Persson, D., Ramløv, H., Westh, P.,<br />

Kristensen, R. M. & Møbjerg, N. (2009b).<br />

Cyclomorphosis in Tardigrada: Adaption to<br />

environmental constraints. Journal of Experimental<br />

Biology, 212:2803-2811.<br />

<strong>Halberg</strong>, K. A. & Møbjerg, N. (2012). First evidence of<br />

epithelial transport in tardigrades: Comparative<br />

investigation of organic anion transport. J. Exp. Biol.,<br />

215:497-507.<br />

<strong>Halberg</strong> K. A., Persson, D., Jørgensen, A. Kristensen,<br />

R. M. and Møbjerg, N. submitted. Population<br />

dynamics of a marine tardigrade: Temperature limits<br />

geographical distribution of Halobiotus crispae.<br />

Mar. Biol. Res.<br />

Hengherr, S., Heyer, A. G., Köhler, H. R. & Schill, R.<br />

(2009). Trehalose and anhydrobiosis in tardigrades –<br />

evidence for divergence in response to dehydration.<br />

FEBS. J., 275:281-288.<br />

Hincha, D. K. & Hagemann, M. (2004). Stabilization<br />

of model membranes during drying by compatible<br />

solutes involved in the stress tolerance of plants and<br />

microorganisms. Biochem. J. 383:277-283.<br />

Hobsen, A. D., Stephenson, W. and Eden A. (1952).<br />

Studies on the physiology of Ascaris lumbricoides:<br />

II. The inorganic composition of the body fluid in<br />

relation to that of the environment. J. Exp. Biol.,<br />

29:22-29.<br />

Hronowski, L. & Armstrong, J. B. (1977). Ion<br />

composition of the plasma of Ambystoma<br />

mexicanum. Comp. Biochem. Physiol. A, 58:181-183.<br />

Jönsson, K. I. & Persson, O. (2010). Trehalose in three<br />

species of desiccation tolerant tardigrades. Open<br />

Zool. J., 3:1-5.<br />

Jørgensen, A., Faurby, S., Hansen, J. G., Møbjerg, N.<br />

and Kristensen, R. M. (2010). Molecular phylogeny<br />

of Arthrotardigrada (Tardigrada). Mol. Phylogen.<br />

Evol., 54:1006-1015.<br />

Normant, M., Kubicka, M., Lapucki, T., Czarnowski,<br />

W. and Michalowska M. (2005). Osmotic and ionic<br />

haemolymph concentration in the Baltic Sea<br />

amphipod Gammarus oceanicus in relation to water<br />

salinity. Comp. Biochem. Physiol. A, 141:94-99.<br />

Macallum, A. B. (1910). The inorganic composition of<br />

the blood in vertebrates and invertebrates, and its<br />

origin. Proc. R. Soc. Lond. B, 82(559):602-624.<br />

Møbjerg N, Dahl C. (1996). Studies on the morphology<br />

and ultrastructure of the Malpighian tubules of<br />

Halobiotus crispae Kristensen, 1982 (Eutardigrada).<br />

Zool. J. Linn. Soc., 116:85-99<br />

Møbjerg, N., A. Jørgensen, J. Eibye-Jacobsen, K. A.<br />

<strong>Halberg</strong>, D. Persson & R. M. Kristensen (2007).<br />

New Records on cyclomorphosis in the marine<br />

eutardigrade Halobiotus crispae (Eutardigrada:<br />

Hypsibiidae). J. Limnol., 66 (suppl. 1): 132-140.<br />

Møbjerg, N. M., <strong>Halberg</strong>, K. A., Persson., D.,<br />

Jørgensen, A. & Kristensen R. M. (2011). Survival<br />

in extreme environments – on current knowledge of<br />

adaptations in tardigrades. Acta Physiol., 202: 409-<br />

420.<br />

Persson, D., <strong>Halberg</strong> K. A., Jørgensen A., Ricci C.,<br />

Møbjerg N. & Kristensen R. M. (2010). Extreme<br />

stress tolerance in tardigrades: Surviving space


conditions in low earth orbit. J. Zool. Syst. Evol.<br />

Res., 49: 90-97.<br />

Persson, D., <strong>Halberg</strong> K. A., Jørgensen A., Møbjerg<br />

N. & Kristensen R. M. (2012). Neuroanatomy of<br />

Halobiotus crispae: Tardigrade brain structure<br />

suggests inclusion into Panarthropoda. J. Morphol.<br />

(in review).<br />

Robertson, J. D. (1949). Ionic regulation in some<br />

marine invertebrates. J. Exp. Biol., 26:182-200.<br />

Robertson, J. D. (1954). The chemical composition of<br />

the blood of some aquatic chordates, including<br />

members of the Tunicata, Cyclostomata and<br />

Osteichthyes. J. Exp. Biol., 31:424-442.<br />

Sands, M, Nicol, S. & McMinn, A. (1998). Fluoride in<br />

Antarctic marine crustaceans. Mar. Biol., 132:591-<br />

598.<br />

Smith R. I. L. (1978). Summer and winter<br />

concentrations of sodium, potassium and calcium in<br />

some maritime Antarctic cryptogams. J. Ecol., 66<br />

(3):891-909.<br />

Sutcliffe, D. W. (1962). The composition of<br />

haemolymph in aquatic insects. J. Exp. Biol.,<br />

39:325-343.<br />

Sutcliffe, D. W. (1963). The chemical composition of<br />

haemolymph in insects and some other arthropods,<br />

in relation to their phylogeny. Comp. Biochem.<br />

Physiol., 9(2):121-135.<br />

Tentori E. and Lockwood, A. P. M. (1990).<br />

Haemolymph magnesium levels in some oceanic<br />

crustaceans. Comp. Biochem. Physiol. A, 4:545-548.<br />

Wilder, M. N., Ikuta, K., Atmomarsono M., Hatta,<br />

T. and Komuro, K. (1998). Changes in osmotic<br />

and ionic concentrations in the hemolymph of<br />

Macrobrachium rosenbergii exposed to varying<br />

salinities and correlation to ionic and crystalline<br />

composition of the cuticle. Comp. Biochem. Physiol.<br />

A, 119:941-950.<br />

Wyatt, G. R. (1961). The biochemistry of insect<br />

hemolymph. Ann. Rev. Entomol., 6:75-102.<br />

Westh, P. and Kristensen, R. M. (1992). Ice formation<br />

in the freeze-tolerant eutardigrades Adorybiotus<br />

coronifer and Amphibolus nebulosus studied by<br />

differential scanning calorimetry. Polar Biol.<br />

12:693-699.<br />

Westh, P. and Ramløv, H. (1991). Trehalose<br />

accumulation in the tardigrade Adorybiotus<br />

coronifer during anhydrobiosis. J. Exp. Zool.<br />

258:303-311.<br />

Yancey, P. H. (2005). Organic osmolytes as<br />

compatible, metabolic and counteracting<br />

cytoprotectants in high osmolarity and other<br />

stresses. J. Exp. Biol., 208:2819-2830.<br />

Inorganic ion composition in Tardigrada


Paper II


497<br />

The Journal of Experimental Biology 215, 497-507<br />

© 2012. Published by The Company of Biologists Ltd<br />

doi:10.1242/jeb.065987<br />

RESEARCH ARTICLE<br />

First evidence of epithelial transport in tardigrades: a comparative investigation of<br />

organic anion transport<br />

<strong>Kenneth</strong> Agerlin <strong>Halberg</strong>* and Nadja Møbjerg<br />

Department of Biology, The August Krogh Building, University of Copenhagen, Universitetsparken 13, DK-2100 Copenhagen Ø,<br />

Denmark<br />

*Author for correspondence (kahalberg@bio.ku.dk)<br />

Accepted 31 October 2011<br />

SUMMARY<br />

We investigated transport of the organic anion Chlorophenol Red (CPR) in the tardigrade Halobiotus crispae using a new method for<br />

quantifying non-fluorescent dyes. We compared the results acquired from the tardigrade with CPR transport data obtained from<br />

Malpighian tubules of the desert locust Schistocerca gregaria. CPR accumulated in the midgut lumen of H. crispae, indicating that<br />

organic anion transport takes place here. Our results show that CPR transport is inhibited by the mitochondrial un-coupler DNP<br />

(1mmoll –1 ; 81% reduction), the Na + /K + -ATPase inhibitor ouabain (10mmoll –1 ; 21% reduction) and the vacuolar H + -ATPase inhibitor<br />

bafilomycin (5mmoll –1 ; 21% reduction), and by the organic anions PAH (10mmoll –1 ; 44% reduction) and probenecid (10mmoll –1 ; 61%<br />

reduction, concentration-dependent inhibition). Transport by locust Malpighian tubules exhibits a similar pharmacological profile,<br />

albeit with markedly higher concentrations of CPR being reached in S. gregaria. Immunolocalization of the Na + /K + -ATPase -subunit<br />

in S. gregaria revealed that this transporter is abundantly expressed and localized to the basal cell membranes. Immunolocalization<br />

data could not be obtained from H. crispae. Our results indicate that organic anion secretion by the tardigrade midgut is transporter<br />

mediated with likely candidates for the basolateral entry step being members of the Oat and/or Oatp transporter families. From our<br />

results, we cautiously suggest that apical H + and possibly basal Na + /K + pumps provide the driving force for the transport; the exact<br />

coupling between electrochemical gradients generated by the pumps and transport of ions, as well as the nature of the apical exit<br />

step, are unknown. This study is, to our knowledge, the first to show active epithelial transport in tardigrades.<br />

Key words: organic anion transport, Chlorophenol Red, 2,4-dinitrophenol, ouabain, bafilomycin, probenecid, para-aminohippuric acid, tardigrade,<br />

insect, V-type H + -ATPase, Na + /K + -ATPase, Malpighian tubule.<br />

INTRODUCTION<br />

The ability to excrete metabolic waste products as well as<br />

environmental toxins (xenobiotics) is a fundamental prerequisite<br />

for animal life. One of the earliest identified systems involved in<br />

such an excretion is the classic organic anion transport system<br />

(Marshall and Vickers, 1923). In vertebrates, this system is<br />

(especially) well known in the proximal tubule where it rids the<br />

organism of various xenobiotic and endobiotic compounds,<br />

through an ATP-dependent, net transepithelial secretory pathway<br />

(reviewed by Dantzler, 2002; Burckardt and Burckardt, 2003;<br />

Russel, 2010). The importance and understanding of similar<br />

transport activities in various invertebrate phyla, however, is<br />

extremely limited. In this study, we expand on the current<br />

knowledge by investigating organic anion transport in the<br />

tardigrade Halobiotus crispae (Tardigrada), as compared with the<br />

desert locust Schistocerca gregaria (Arthropoda).<br />

The phylum Tardigrada consists of a group of minute, eight-legged,<br />

multicellular animals that, like arthropods and nematodes, belong to<br />

the invertebrate superclade Ecdyzosoa (Aguinaldo et al., 1997). They<br />

are considered essential to our understanding of early metazoan<br />

evolution, yet fundamental questions concerning their basic biology<br />

remain unanswered (Møbjerg et al., 2011). In spite of their small size<br />

(~50–1200mm), tardigrades are relatively complex animals; they are<br />

composed of >1000 cells, and possess a well-developed musculature<br />

and nervous system, as well as a complex alimentary canal and<br />

specialized reproductive and excretory organs (Dewel and Dewel,<br />

1979; Rebecchi and Bertolani, 1994; Møbjerg and Dahl, 1996;<br />

Schmidt-Raesa and Kulessa, 2007; Pelzer et al., 2007; Zantke et al.,<br />

2008; <strong>Halberg</strong> et al., 2009a; Møbjerg et al., 2011; Rost-Roszkowska<br />

et al., 2011). The alimentary canal can be divided into several<br />

morphologically distinct regions i.e. bucco-pharyngeal apparatus,<br />

oesophagus, midgut and hindgut, with the Malpighian tubules (MTs)<br />

of eutardigrades positioned at the junction between these last two<br />

sections. Interestingly, the same basic organizational pattern is found<br />

in insects, and has been used as a strong argument for the homology<br />

of these two organ systems (Greven, 1982; Møbjerg and Dahl, 1996).<br />

Among multi-cellular animals, tardigrades exhibit an extraordinary<br />

ability to resist environmental extremes, and are known to survive<br />

conditions greatly exceeding those encountered in their natural habitat<br />

– even in space (Jönsson et al., 2008; Rebecchi et al., 2008; Persson<br />

et al., 2011). The biochemical and physiological mechanisms<br />

mediating this unique tolerance, however, remain largely unidentified.<br />

Previously, we have shown that the marine eutardigrade H. crispae<br />

tolerates large changes in external salinity surviving periods of osmotic<br />

stress by maintaining haemolymph osmotic pressure above that of<br />

the external medium (<strong>Halberg</strong> et al., 2009b); an adaptive mechanism<br />

likely present in all eutardigrades (Møbjerg et al., 2011). Here, we<br />

identify organs involved in transepithelial transport of organic anions<br />

and investigate the transport characteristics with the aim of providing<br />

a better understanding of the unique biology in these animals.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


498<br />

K. A. <strong>Halberg</strong> and N. Møbjerg<br />

Organic anion transport has previously been described in other<br />

groups of invertebrates (George et al., 1999; Torrie et al., 2004;<br />

Faria et al., 2010). Notably, the alimentary canal and MTs of insects,<br />

which collectively form the functional analogue of the vertebrate<br />

kidney, have been studied (reviewed by Phillips, 1981; O’Donnell<br />

et al., 2003; Dow and Davies, 2006). As in vertebrates, the excretory<br />

organs of insects transport a wide range of organic solutes and<br />

exogenous toxins through organic anion transporters (Oats), organic<br />

anion-transporting peptides (Oatps), P-glycoproteins (Mdr/P-gp) as<br />

well as multidrug resistance-associated proteins (Mrps) (Maddrell<br />

et al., 1974; Bresler et al., 1990; Lanning et al., 1996; Linton and<br />

O’Donnell, 2000; Torrie et al., 2004; Neufeld et al., 2005; Leader<br />

and O’Donnell, 2005; O’Donnell and Leader, 2006; Chahine and<br />

O’Donnell, 2009; Chahine and O’Donnell, 2010). Organic anions<br />

are divided into type I and type II organic anions, on the basis of<br />

structural and chemical properties (e.g. Wright and Dantzler, 2004).<br />

The different groups of transporters vary in their transport<br />

mechanisms, but overlap in their substrate specificity, as they<br />

transport carboxylates and sulphonates interchangeably (Neufeld et<br />

al., 2005; Chahine and O’Donnell, 2009). Oats and Oatps (solute<br />

carriers belonging to the SLC22 and SLC21/SLCO family) transport<br />

both small (450Da) hydrophobic type II organic anions (Oatps),<br />

whereas Mdr/P-gp and Mrps (ABC transporters, ABCB and ABCC<br />

subfamilies) generally transport large (>500Da) polyvalent type II<br />

organic anions (Wright and Dantzler, 2004; Russel, 2010). From<br />

the insect’s perspective, the clearance of exogenous toxins is of<br />

particular interest, as insects often live in environments with high<br />

xenobiotic exposure, potentially at harmful or lethal concentrations.<br />

Accordingly, they are forced to process naturally occurring plant<br />

toxins (Torrie et al., 2004; Neufeld et al., 2005), as well as<br />

anthropogenic contaminants, such as insecticides (Lanning et al.,<br />

1996; Neufeld et al., 2005; Buss and Callaghan, 2008). The<br />

physiological importance of this detoxification system – and<br />

implicitly the MTs – is emphasized by the fact that transcripts for<br />

these transporters are enriched in the transcriptome of the MTs<br />

(Wang et al., 2004), and because the expression of several Oatp and<br />

Mdr transporters is significantly upregulated upon dietary exposure<br />

to Oatp and Mdr substrates (Mulenga et al., 2008; Chahine and<br />

O’Donnell, 2009; Chahine and O’Donnell, 2010). Indeed, it has been<br />

suggested that organic solute excretion is the most significant<br />

function of the insect MT (Dow and Davies, 2006). Nevertheless,<br />

in spite of significant efforts in the past decades, the mechanisms<br />

underlying transepithelial transport of organic anions is far from<br />

understood, and our knowledge is limited to relatively few taxa.<br />

In this study we examined epithelial transport in tardigrades.<br />

Using a comparative approach, we investigated the sites,<br />

characteristics and pharmacological profile of the net transepithelial<br />

transport of Chlorophenol Red (CPR; 3,3-dichlorophenolsulphone-phthalein),<br />

a pH indicator and a prototypic substrate of<br />

the classic organic anion secretory pathway, in the tardigrade H.<br />

crispae and the desert locust S. gregaria. Our results show that the<br />

tardigrade midgut is the principal site of CPR transport, and that<br />

this transport is active and transporter mediated. Additionally, our<br />

data show that the pharmacological profiles of CPR transport in the<br />

tardigrade midgut and locust MT are surprisingly similar.<br />

MATERIALS AND METHODS<br />

Quantification of CPR accumulation<br />

The quantification of CPR accumulation was performed by<br />

introducing a new method for quantifying non-fluorescent dyes. The<br />

method exploits the optical properties of CPR by relating relative<br />

Difference in red spectrum/green spectrum<br />

relationship (%)<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

0<br />

pH 6<br />

r 2 =0.994<br />

pH 7–10<br />

r 2 =0.987<br />

0 1 2 3 4 5<br />

[CPR] (mmol l –1 )<br />

pH 11<br />

r 2 =0.996<br />

Fig.1. Standard curves showing the percentage difference in the red<br />

spectrum/green spectrum relationship as a function of Chlorophenol Red<br />

(CPR) concentration at different pH values (pH6, 7–10 and 11). The curves<br />

were fitted by regression analysis using OriginPro 7.5 with a third order<br />

polynomial providing the best fit with r 2 values close to unity.<br />

changes in spectral light to that of dye concentration (Fig.1). As<br />

CPR exhibits a pH-dependent shift in colour (from yellow to violet)<br />

between pH4.6 and 7.0, and a secondary shift (to purple) at pH>10,<br />

a standard curve was constructed at three discrete colours relevant<br />

to our investigation (red, violet and purple corresponding to<br />

pH6–11). This was done in order to correct for potential pH<br />

dependent effects on CPR quantification caused by the tissue. The<br />

standard curves were constructed from optical analysis of<br />

micrographs taken of samples (40ml) with known dye concentrations<br />

(ranging from 0.1 to 5mmoll –1 ) at each respective pH value. The<br />

optical measurements provided relative, arbitrary values for the light<br />

intensities of the red, green and blue spectrum of spectral light from<br />

each sample, and were acquired using the image analysis and<br />

visualization software Imaris 6.4 (Bitplane, Zurich, Switzerland).<br />

The standard curves are expressed as percentage difference in the<br />

red spectrum/green spectrum relationship as a function of CPR<br />

concentration (Fig.1). The curves were fitted by regression analysis<br />

using OriginPro 7.5 (OriginLab, Northampton, MA, USA) with a<br />

third order polynomial providing the best fit with r 2 values close to<br />

unity. The CPR concentration in a given tissue was subsequently<br />

calculated by employing the appropriate standard curve. The<br />

standard curve was chosen by visually comparing the colour of the<br />

accumulated CPR with that of the respective standard curves.<br />

Contrary to our expectations, the midgut of H. crispae was the<br />

only organ in which CPR accumulation was clearly visualized.<br />

No accumulation was observed in the tardigrade MTs. As such,<br />

quantification of dye accumulation was only performed from the<br />

tardigrade midgut, and compared with the accumulation in the<br />

MTs of S. gregaria. Specifically, 6–12 regions from areas of the<br />

tardigrade midgut containing the highest dye intensity were<br />

selected arbitrarily, in addition to a similar number from the<br />

adjacent haemolymph, while four different regions were selected<br />

from each insect MT. Each region selected represented a circle<br />

with a diameter of 5mm for H. crispae and 25mm for S. gregaria.<br />

The average light intensity within these circles was measured by<br />

the Imaris program. Averaging the measured intensity of all<br />

selected regions provided an overall average intensity for the red,<br />

green and blue colour spectrum within each investigated organ<br />

following each exposure. As CPR predominantly appears red (to<br />

violet) in the midgut of H. crispae and in the MTs of S. gregaria,<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Epithelial transport in tardigrades<br />

499<br />

the red spectrum/green spectrum relationship offered an estimate<br />

of dye accumulation, as calculated from the appropriate standard<br />

curve. The final CPR concentration was normalized according to<br />

the background light intensity, i.e. the tardigrade haemolymph or<br />

untreated (control) insect MTs. For tardigrades, only regions<br />

without gut content were selected for quantification, in order to<br />

avoid the influence of gut content on the wavelength of captured<br />

light. Deviations in animal depth did not influence CPR<br />

quantification within the variation encountered in this study. The<br />

MTs of S. gregaria consist of three structurally distinct regions,<br />

i.e. proximal, middle and distal relative to the gut (Garret et al.,<br />

1988). Dye accumulation was estimated from the proximal region<br />

and parts of the middle region within 5mm of the junction with<br />

the gut, as dye accumulation was highest here.<br />

Test solutions<br />

The haemolymph osmolality of H. crispae, kept at a salinity of<br />

20p.p.t., was previously measured by nanolitre osmometry to<br />

~950mOsmkg –1 (<strong>Halberg</strong> et al., 2009b). At present we do not know<br />

the composition of tardigrade extracellular fluids. As such, the<br />

experimental solution (control solution) was prepared from<br />

evaporative reduction of seawater (SW; salinity 20p.p.t., pH8)<br />

collected at the locality, and 35mmoll –1 glucose was added to<br />

alleviate potential variation in experiments caused by differences<br />

in nutrient availability. This yielded a final measured osmolality of<br />

950±3mOsmkg –1 (N3). For S. gregaria an insect saline (control<br />

solution) was prepared containing (in mmoll –1 ): 130 NaCl, 10 KCl,<br />

4 NaHCO 3 , 2 MgSO 4 , 2 CaCl 2 , 6 NaH 2 PO 4 , 35 glucose and 5 Hepes,<br />

titrated to pH7.2, with a measured osmolality of 336±2mOsmkg –1<br />

(N3).<br />

In order to explore the kinetics of CPR transport by MTs of S.<br />

gregaria, a concentration–response curve was constructed over a<br />

5000-fold range of CPR concentrations (1mmoll –1 to 5mmoll –1 )<br />

(Fig.2). These data revealed that CPR transport is saturated at an<br />

external CPR concentration of ~1.6mmoll –1 (V max 1.58mmoll –1 ,<br />

K m 81.8mmoll –1 ; Fig.2). However, the MTs are unable to<br />

concentrate the dye at high concentrations of CPR (>1mmoll –1 ;<br />

Fig.2, inset). Experiments, on both animals, were performed at a<br />

concentration of 1mmoll –1 CPR.<br />

Test solutions were prepared from the two control solutions<br />

containing 1mmoll –1 CPR, and one of the following inhibitors: 2,4-<br />

dinitrophenol (DNP, 1mmoll –1 ), ouabain (10mmoll –1 ), bafilomycin<br />

A 1 (5mmoll –1 ), para-aminohippuric acid (PAH; 10mmoll –1 ) or<br />

probenecid (0.1–10mmoll –1 ). The final osmolality of the solutions<br />

was measured on a Vapro 5520 vapour pressure osmometer (Wescor,<br />

Logan, UT, USA). All solutions were titrated with NaOH to pH8<br />

for H. crispae and pH7.2 for S. gregaria.<br />

Experimental animals<br />

Specimens of H. crispae Kristensen 1982 were collected on 11<br />

February 2008 and 17 January 2010 at Vellerup Vig, Isefjord,<br />

Denmark (55°44.206N, 11°51.258E) (see Fig.3A). Bottom samples<br />

were collected with a mini van Veen grab at a depth of 1–2m (salinity<br />

~20p.p.t., pH8). Rocks, algae and sediment collected with the grab<br />

were freshwater shocked. The debris was decanted into a conical net<br />

(mesh size 63mm) and subsequently transferred to SW from the<br />

locality and kept at 4°C. Animals in the active stage (see Kristensen,<br />

1982; Møbjerg et al., 2007; <strong>Halberg</strong> et al., 2009b) were identified<br />

using a dissection microscope. The tardigrades were kept for a period<br />

of up to 6months at 4°C in SW (salinity 20p.p.t., pH8) and regularly<br />

supplied with fresh substrate, consisting mainly of sediment, organic<br />

debris, filamentous algae and diatoms.<br />

A<br />

B<br />

[CPR]lumen (mmol l –1 )<br />

2.2<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

tc<br />

1 μmol l –1 10 μmol l –1 100 μmol l –1 1 mmol l –1 5 mmol l –1<br />

lu<br />

tc<br />

lu<br />

K m =81.8 μmol l –1<br />

V max =1.58 mmol l –1<br />

Lumen/bath ratio<br />

tc<br />

200<br />

150<br />

100<br />

Specimens of S. gregaria Forskål 1775 were acquired from a<br />

specialized animal shop (www.exopark.dk), and kept at room<br />

temperature (RT) with light:dark periods of 16h:8h and fed annual<br />

meadow grass (Poa annua) ad libitum.<br />

Exposure to test solutions<br />

Halobiotus crispae<br />

Evaluation of CPR transport by tardigrade epithelia was performed<br />

on whole animals immersed in the dye solution. Data obtained from<br />

82 animals were used for the study – no distinction was made<br />

between male and female specimens. Initial observations revealed<br />

that the animals did not take up dye through the mouth or cuticle<br />

(CPR non-punctured, Fig.3) and test solutions were therefore<br />

introduced to the haemolymph of single specimens (length<br />

300–500mm) through a small hole made in the cuticle in the anterior<br />

part of the animal. The animal was incubated for a period of 60min<br />

at RT in the respective test solution, and quickly washed in SW<br />

prior to photography. Light micrographs of the specimens were taken<br />

in bright-field at a 40 magnification, using an Olympus DP20<br />

camera mounted on an Olympus BX50 microscope (Olympus,<br />

Hamburg, Germany). Additionally, in order to investigate whether<br />

the test solutions were ingested during the experimental period, intact<br />

non-punctured animals were incubated in the CPR test solution for<br />

a corresponding period. The animals were washed and photographed<br />

as described above.<br />

50<br />

lu<br />

tc<br />

[CPR] bath (mmol l –1 )<br />

lu<br />

0<br />

0 0.5 1<br />

0 1 2 3 4 5<br />

Fig.2. Accumulation of CPR as a function of external CPR concentration in<br />

Malpighian tubules (MTs) of Schistocerca gregaria. (A)Representative light<br />

micrographs of the MTs following exposure to different concentrations<br />

(1mmoll –1 , 10mmoll –1 , 100mmoll –1 , 1mmoll –1 and 5mmoll –1 ) of CPR.<br />

Scale bars, 100mm. lu, lumen; tc, trachea. (B)Luminal CPR concentration<br />

as a function of bath CPR concentration, revealing the kinetic parameters<br />

K m and V max for CPR transport. Each point shows the mean ± s.d. for<br />

N4–6 animals with 3–5 MTs providing the estimate for each animal. The<br />

solid line represents the fit to the Michaelis–Menten equation by non-linear<br />

regression analysis (using error as weight). Insert shows the lumen/bath<br />

ratio of CPR as a function of external CPR concentration.<br />

lu<br />

5<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


500<br />

K. A. <strong>Halberg</strong> and N. Møbjerg<br />

B<br />

[CPR]luminal (mmol l –1 )<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

SW<br />

control<br />

mg<br />

C<br />

(12)<br />

SW control<br />

CPR nonpunctured<br />

mg<br />

(6)<br />

CPR non-punctured<br />

mg<br />

CPR<br />

(13)<br />

1 mmol l –1 CPR<br />

+DNP<br />

–81%<br />

mg<br />

**<br />

(5)<br />

+1 mmol l –1 DNP<br />

+Bafilomycin<br />

+Ouabain +DMSO<br />

–21% –21%<br />

n n<br />

mg<br />

*<br />

+10 mmol l –1 ouabain<br />

mg<br />

*<br />

(6) (8)<br />

+5 μmol l –1 bafilomycin<br />

A<br />

+PAH<br />

–44%<br />

mg<br />

*<br />

*<br />

**<br />

(11)<br />

+10 mmol l –1 PAH<br />

mg<br />

*<br />

Leg 4<br />

+Probenecid<br />

–61%<br />

mg<br />

Leg 1<br />

Leg 2<br />

Leg 3<br />

**<br />

(10)<br />

+10 mmol l –1 probenecid<br />

Fig.3. CPR accumulation in the gut lumen of Halobiotus crispae,<br />

and the effect of inhibitors on dye accumulation. (A)Light<br />

micrograph of H. crispae from a ventral view showing the animalʼs<br />

basic morphology. mg, midgut. Asterisks indicate MTs. Scale bar,<br />

100mm. (B)Representative light micrographs of the midgut (mg)<br />

following exposure to the various test solutions. Seawater control<br />

shows the midgut of a punctured animal following immersion in<br />

seawater, while CPR non-punctured shows the midgut of an intact<br />

animal after immersion in CPR solutions. Estimations of luminal<br />

CPR concentration were performed solely from areas devoid of gut<br />

content, i.e. brown material in the gut. The percentage change in<br />

CPR concentration, compared with experiments on CPR alone, is<br />

noted. Scale bars, 50mm. (C)Corresponding luminal concentrations<br />

of CPR. Data are depicted as means ± s.d. Asterisks refer to a<br />

significant difference from CPR alone (*P


Epithelial transport in tardigrades<br />

501<br />

Microsystems, Wetzlar, Germany) with glass knives and<br />

subsequently stained with Toluidine Blue.<br />

For immunocytochemistry, whole animals of H. crispae and MTs<br />

of S. gregaria were fixed in 3% paraformaldehyde in 0.1moll –1<br />

sodium cacodylate buffer (pH7.4) for 60min and subsequently<br />

transferred to 0.1moll –1 sodium cacodylate buffer. The tissue was<br />

then dehydrated through a graded series of ethanol and xylene,<br />

embedded in paraffin and sectioned into ~10mm sections, or<br />

transferred to PBS and used as whole mounts for immunostaining.<br />

Paraffin sections were deparaffinized through a graded series of xylene<br />

and alcohol, washed in saline (control solution; see ‘Test solutions’<br />

above) and blocked with 10% normal goat serum (Invitrogen,<br />

Carlsbad, CA, USA) for 30min, prior to incubation with primary<br />

antibody. Paraffin sections, as well as whole mounts, were incubated<br />

overnight at 4°C in insect saline (MTs) or PBS (tardigrades) containing<br />

10% normal goat serum, 0.1% Triton-X and primary antibody. The<br />

Na + /K + -ATPase -subunit monoclonal mouse primary antibody 5-<br />

IgG (10mgml –1 ) was developed by D. M. Famborough, and obtained<br />

from the Developmental Studies Hybridoma Bank (University of<br />

Iowa, Iowa City, IA, USA). This antibody has been used to identify<br />

the Na + /K + -ATPase in numerous excretory tissues, including MTs of<br />

Drosophila melanogaster (Lebovitz et al., 1989; Torrie et al., 2004),<br />

the gills of the blue crab Callinectes sapidus (Towle et al., 2001) and<br />

the pronephros of Ambystoma mexicanum (Haugen et al., 2010).<br />

Following an extensive wash in saline, the tissue was incubated with<br />

(anti-mouse) Alexa Fluor 594 (1:100) secondary antibody (Invitrogen)<br />

overnight at 4°C. The tissue was rinsed with saline then counterstained<br />

with Alexa Fluor 488-conjugated phalloidin (luminal marker; 1:40;<br />

Invitrogen) and DAPI (50mgml –1 ; Invitrogen) for 2h, washed and<br />

mounted on glass coverslips in Vectashield (Vector Laboratories Inc.,<br />

Burlingame, CA, USA). Images were acquired using a Leica DM<br />

RXE 6 TL microscope equipped with a Leica TCS SP2 AOBS<br />

confocal laser scanning unit (Leica Microsystems, Wetzlar, Germany).<br />

The tissue was scanned employing sequential scanning (setting:<br />

between frames) using the 488nm line of an argon/krypton laser and<br />

the 594nm line of a helium laser, in addition to the 405nm UV laser<br />

line. The image series was processed and edited using Imaris software.<br />

Confocal images are based on 240 optical sections of a Z-series<br />

performed at intervals of 0.7mm. Experiments were conducted<br />

multiple times with corresponding results. All control preparations<br />

without primary antibody were negative for immunostaining.<br />

Chemicals<br />

All chemicals were obtained from Sigma-Aldrich (St Louis, MO,<br />

USA). Bafilomycin was dissolved and stored in dimethyl sulphoxide<br />

(DMSO). Inhibitors were allowed to dissolve in the CPR solution<br />

overnight prior to use.<br />

Statistics<br />

Data are expressed as means ± s.d. unless otherwise stated. The<br />

statistical significance of differences between the various exposures<br />

was tested using one-way ANOVA followed by a Tukey multiple<br />

comparisons of means. The statistical tests were performed using<br />

the data analysis program OriginPro 7.5 (OriginLab). Significance<br />

levels were P>0.05 (not significant), P


502<br />

K. A. <strong>Halberg</strong> and N. Møbjerg<br />

[CPR]lumen (mmol l –1 )<br />

A<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

CPR<br />

mg<br />

B<br />

(13)<br />

*<br />

+Bafilomycin<br />

+DMSO<br />

–14%<br />

+DMSO<br />

–21%<br />

CPR<br />

mg<br />

**<br />

*<br />

gm<br />

(5) (8)<br />

C<br />

tc lu lu tc lu<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

D<br />

+DMSO<br />

+5%<br />

**<br />

(11) (10)<br />

+Bafilomycin<br />

+DMSO<br />

–26%<br />

**<br />

(8)<br />

tc<br />

Fig.6. The effect of bafilomycin and dimethylsulphoxide (DMSO) on<br />

CPR accumulation. (A)Representative light micrographs of the<br />

tardigrade midgut (mg) following exposure to the test solutions.<br />

The percentage change in CPR concentration, compared with<br />

experiments with CPR alone, is noted. Scale bars, 50mm.<br />

(B)Corresponding luminal concentration of CPR in midgut lumen.<br />

(C)Representative light micrographs of locust MTs following<br />

exposure to the test solutions. The percentage change in CPR<br />

concentration, compared with experiments on CPR alone, is noted.<br />

Scale bars, 100mm. (D)Corresponding luminal concentration of<br />

CPR in the MTs. Asterisks refer to significant difference from<br />

1mmoll –1 CPR alone (*P


Epithelial transport in tardigrades<br />

503<br />

A<br />

[CPR]lumen (mmol l –1 )<br />

1.6<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

+0.1 mmol l –1<br />

CPR –26%<br />

CPR<br />

mg<br />

B<br />

(13)<br />

**<br />

mg<br />

** **<br />

(7)<br />

+1 mmol l –1<br />

–56%<br />

*<br />

mg<br />

**<br />

+10 mmol l –1<br />

–61%<br />

mg<br />

(4) (10)<br />

C<br />

tc<br />

lu<br />

lu<br />

1.6 D ** **<br />

(11) **<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

CPR<br />

+0.1 mmol l –1<br />

–9%<br />

(5)<br />

+1 mmol l –1<br />

–47%<br />

lu<br />

(5)<br />

+10 mmol l –1<br />

–77%<br />

tc<br />

**<br />

lu<br />

(9)<br />

Fig.7. Titration of the inhibitory effect of<br />

probenecid. (A)Representative light<br />

micrographs of the tardigrade midgut (mg)<br />

following exposure to the different<br />

concentrations (0.1, 1 and 10mmoll –1 ) of<br />

probenecid. The percentage change in CPR<br />

concentration, compared with experiments<br />

with CPR alone, is noted. Scale bars, 50mm.<br />

(B)Corresponding luminal concentration of<br />

CPR in midgut lumen. (C)Representative<br />

light micrographs of the insect MTs following<br />

exposure to the different concentrations (0.1,<br />

1 and 10mmoll –1 ) of probenecid. The<br />

percentage change in CPR concentration,<br />

compared with experiments on CPR alone, is<br />

noted. Scale bars, 100mm. (D)Corresponding<br />

luminal concentration of CPR in the MTs.<br />

Asterisks refer to significant difference from<br />

1mmoll –1 CPR alone (*P


504<br />

K. A. <strong>Halberg</strong> and N. Møbjerg<br />

probenecid concentration and CPR accumulation (and activity; data<br />

not shown). At a concentration of 0.1mmoll –1 probenecid, CPR<br />

concentration was reduced by 26% with a luminal concentration of<br />

0.42±0.07mmoll –1 in H. crispae, and by 9% corresponding to a<br />

concentration of 1.2±0.14mmoll –1 in S. gregaria. When applied at<br />

a concentration of 1mmoll –1 , probenecid reduced CPR concentration<br />

by 56% corresponding to a luminal concentration of<br />

0.25±0.05mmoll –1 in H. crispae, and by 47% equivalent to<br />

0.7±0.16mmoll –1 in S. gregaria (Fig.7).<br />

Midgut and MT structure and immunostaining<br />

Our transport studies indicate that CPR transport is active and<br />

transporter mediated in the midgut of H. crispae as well as the MTs<br />

of S. gregaria. It is well established that the vacuolar H + -ATPase<br />

is important for energizing transepithelial transport in insects,<br />

whereas (and in contrast to vertebrate literature) the role of the<br />

Na + /K + -ATPase is controversial (e.g. Torrie et al., 2004; Beyenbach<br />

et al., 2010). The data obtained in the present study using ouabain<br />

would suggest that the Na + /K + -ATPase is important for CPR<br />

transport in both animals. Conversely, ouabain could act as a<br />

competitive inhibitor, being transported by the same transporters as<br />

CPR. We therefore investigated whether the Na + /K + pump is<br />

expressed in the epithelia.<br />

Unfortunately, our attempt to localize Na + /K + -ATPase in<br />

tardigrades was unsuccessful, both on whole mounts and on paraffin<br />

sections, and also in attempts at antigen retrieval.<br />

Our observations on the MTs of S. gregaria generally correspond<br />

with those previously described (Garret et al., 1988), when focusing<br />

on the proximal region and parts of the middle region used in the<br />

present study. In general, we observed a prominent brushborder<br />

(Fig.8B,C) and at least two cell types. One type, a principal-like cell,<br />

has a single nucleus, while the other is a very large double-nucleated<br />

cell (dashed circles in Fig.8). Cross-striated muscle fibers are present<br />

in both the transverse (Fig.8A,B) and longitudinal direction<br />

(Fig.8E,G). The transverse muscle fibers were only observed in the<br />

proximal-most region of the MTs, whereas the longitudinal muscle<br />

extended the entire length of the tubule. Visual observations revealed<br />

that the transverse muscles aid the clearance of luminal content into<br />

the midgut through peristaltic contractions. Our immunocytochemical<br />

analysis shows that the Na + /K + -ATPase is expressed in the MTs of<br />

S. gregaria, and is localized to the basal plasma membranes<br />

(Fig.8D–G), although to a lesser extent in the double-nucleated cells<br />

(Fig.8D,G). The localization was confirmed by counterstaining of<br />

nuclei by DAPI and staining apical microvilli with phalloidin. This<br />

is, to our knowledge, the first study showing immunolocalization of<br />

the Na + /K + -ATPase in S. gregaria MTs. Interestingly, the localization<br />

of the transporter substantiates previously published data on isolated<br />

basal membrane fractions from MTs and hindgut from S. gregaria,<br />

which show a prominent Na + /K + -ATPase activity (Al-Fifi, 2007).<br />

DISCUSSION<br />

Our study is significant in two main aspects: (i) we are the first to<br />

provide evidence for active epithelial transport in tardigrades, and<br />

(ii) our data show that tardigrades possess an organic anion transport<br />

system. Consequently, our study is important for understanding the<br />

evolution of transport systems.<br />

CPR is categorized as a sulphonate, but is known to compete with<br />

substrates for both solute carriers (SLC22 and SLC21/SLCO) and<br />

ABC transporters in a number of organisms (Pritchard et al., 1999;<br />

Linton and O’Donnell, 2000; Chahine and O’Donnell, 2009). In H.<br />

crispae, CPR transport is ATP dependent (strong inhibition by DNP),<br />

probably energized by both the Na + /K + -ATPase and the V-type H + -<br />

ATPase (bafilomycin and ouabain reduce CPR accumulation), and<br />

possibly transporter mediated (inhibited by the prototypic organic<br />

anions PAH and probenecid). The latter suggests that CPR transport<br />

is (at least partly) transcellular. Comparing the transport characteristics<br />

of CPR transport between the tardigrade midgut and the insect MT<br />

reveals a surprisingly similar overall pharmacological profile of the<br />

investigated tissues, albeit with markedly higher concentrations of<br />

CPR observed in the insect MT. In addition to transcellular transport<br />

of the dye, fluid secretion may augment transport of organic anions<br />

by convective secretion through the paracellular pathway, and/or<br />

reduce diffusive back-flux of organic anions from the tubule lumen<br />

to the haemolymph (O’Donnell and Leader, 2006; Chahine and<br />

O’Donnell, 2010). Consequently, without a measure of fluid secretion<br />

rates, and thereby a measure of net CPR secreted, we cannot conclude<br />

whether the difference in relative concentration of CPR in the gut<br />

lumen of H. crispae, compared with the MTs of S. gregaria, is a<br />

consequence of a lower dye transport capacity, or whether it reflects<br />

differences in fluid transport rates. Also, we cannot be certain that<br />

the concentration of CPR in the haemolymph of H. crispae is exactly<br />

the same as that of the surrounding bath – although, the relatively<br />

high dye concentration used along with the small diffusion distances<br />

involved, the relatively long exposure time (60min) and the animal<br />

movements that facilitate fluid exchange, would make potential<br />

concentration differences negligible.<br />

A vast number of papers have investigated various aspects of<br />

organic anion transport in insect MTs. In the present study we<br />

observed a mean CPR concentration of 1.3±0.13mmoll –1 when no<br />

inhibitors were added. This concentration is elevated above the bath<br />

concentration (1mmoll –1 ), albeit by a much smaller factor than that<br />

reported from other insects (Maddrell et al., 1974; Bresler et al.,<br />

1990; Linton and O’Donnell, 2000; Leader and O’Donnell, 2005).<br />

However, when exposed to external CPR concentrations of 1, 10<br />

and 100mmoll –1 , the mean luminal CPR concentrations were<br />

0.16±0.03, 0.73±0.09 and 0.89±0.07mmoll –1 , respectively, which<br />

is a factor of ~160, 70 and 9 above bath concentrations (Fig.2).<br />

When exposed to a bath concentration of 5mmoll –1 CPR, the mean<br />

luminal concentration was 1.79±0.33mmoll –1 , a factor of 2.8 below<br />

external concentrations. Consequently, the luminal CPR<br />

concentration is maximally elevated (~160-fold) above that in the<br />

bathing medium when the latter contains CPR at a concentration of<br />

1mmoll –1 . This concentrative ability is among the highest measured<br />

for insects, and is an additional confirmation of active transport of<br />

CPR by the insect MTs.<br />

Ouabain, a well-characterized, potent inhibitor of the Na + /K + -<br />

ATPase, reduced the CPR concentration in both the tardigrade<br />

midgut and the insect MTs by ~20–21%. A similar 23% reduction<br />

in PAH secretion in the presence of 1mmoll –1 ouabain was reported<br />

from MTs of D. melanogaster (Linton and O’Donnell, 2000), while<br />

fluorescein uptake was reduced in MTs of Blaberus giganteus by<br />

30% and in MTs of Locusta migratoria by 20%, when ouabain was<br />

used at 1 and 0.1mmoll –1 , respectively (Bresler et al., 1990). Our<br />

immunocytochemical investigation on the MTs of S. gregaria<br />

revealed expression of the Na + /K + -ATPase in the basal cell<br />

membranes, although to a lesser extent in the double-nucleated cells.<br />

A basal localization of the Na + /K + -ATPase in principal cells of MTs<br />

of D. melanogaster has previously been reported (Torrie et al., 2004).<br />

The fact that there is expression of the Na + /K + -ATPase in the basal<br />

membranes, and that CPR transport is ouabain sensitive, suggests<br />

that the pump is important for CPR transport. Alternatively, it could<br />

be argued that the inhibitory effects of ouabain are due to competitive<br />

inhibition, rather than to non-competitive inhibition, given the fact<br />

that ouabain is actively transported by members of the SLC21/SLCO<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Epithelial transport in tardigrades<br />

505<br />

subfamily (Oatps) in D. melanogaster (Torrie et al., 2004). Whether<br />

this is ubiquitous in insects is at present not known; however, several<br />

members of the Oatp family are known to transport ouabain in<br />

human tissue (see Hagenbuch and Gui, 2008). Competition studies<br />

using fluorescently labeled ouabain would help in clarifying this<br />

matter. Considering the striking similarities between the<br />

pharmacological profiles of the tardigrade midgut and insect MT,<br />

it seems reasonable to assume that the Na + /K + -ATPase is similarly<br />

present in the basal cell membranes of the tardigrade midgut (in<br />

spite of failed attempts to localize this transporter). In support of<br />

this interpretation is the fact that transcripts for the -subunit of the<br />

Na + /K + -ATPase were found in the expressed sequence tags (EST)<br />

library from Hybsibius dujardini Doyère, 1840 (TardiBASE cluster<br />

ID: HDC01733 TardiBASE; http://xyala.cap.ed.ac.uk/research/<br />

tardigrades/tardibase.shtml).<br />

Bafilomycin is a specific inhibitor of the V-type H + -ATPase and<br />

was found to reduce CPR accumulation in both H. crispae and S.<br />

gregaria. This observation is consistent with the fact that the V-type<br />

H + -ATPase is viewed as being central to the transport activities of<br />

the MT in insects (Weng et al., 2003; Beyenbach et al., 2010) and<br />

perhaps also for transport in tardigrades; the B-, C-, D-, E-, G- and<br />

H-subunits of the V-type H + -ATPase were found in the EST library<br />

of H. dujardini (TardiBASE). Indeed, visual changes to CPR colour<br />

were observed in the midgut of H. crispae (CPR became purple; data<br />

not shown) 10min post-incubation, suggesting an alkalization of the<br />

midgut content. This observation is in accordance with specific<br />

inhibition of an apical V-type H + -ATPase in the tardigrade midgut.<br />

PAH is a prototypical substrate of the classic organic anion transport<br />

system (i.e. the SLC22 subfamily) and was shown to be an effective<br />

inhibitor of CPR transport in both H. crispae and S. gregaria (~40%<br />

reduction in both animals). This finding is a strong indication that<br />

PAH (carboxylate) and CPR (sulphonate) transport are mediated by<br />

a common transporter, or alternatively through two separate transport<br />

systems overlapping in affinity, in both investigated epithelia. Whether<br />

carboxylates (e.g. PAH and probenecid) and sulphonates (e.g. CPR)<br />

are handled by a common transporter (Bresler et al., 1990), or by two<br />

separate transport systems (Maddrell et al., 1974; Linton and<br />

O’Donnell, 2000; Chahine and O’Donnell, 2009), appears to be highly<br />

variable and/or species specific. Indeed, transport of PAH and<br />

probenecid (carboxylates) and CPR and methotrexate (MTX;<br />

sulphonates) are all mediated by one tranporter (OAT1) in humans,<br />

but not in rats (Burckardt and Burckardt, 2003), whereas MTX uptake<br />

is competitively inhibited by CPR and probenecid, but not PAH in<br />

D. melongaster (Cahine and O’Donnell, 2009).<br />

Probenecid is implicitly regarded as a competitive inhibitor of<br />

organic anion transport, and has been shown to compete with both<br />

solute carrier and ABC transporter substrates (Bresler et al., 1990;<br />

Linton and O’Donnell, 2000; Neufeld et al., 2005; Leader and<br />

O’Donnell, 2005). It is typically applied in a concentration of<br />

1mmoll –1 ; here, we examined the effects of probenecid on CPR<br />

transport in the range 0.1–10mmoll –1 . In H. crispae, probenecid<br />

reduced CPR accumulation by 26%, 56% and 61% at a concentration<br />

of 0.1, 1 and 10mmoll –1 , respectively, whereas the corresponding<br />

reduction in CPR concentration was 9%, 47% and 77% in S.<br />

gregaria. In the presence of very high probenecid concentrations<br />

(10mmoll –1 ), we observed an almost total loss of motility in H.<br />

crispae. Indeed, the level of animal movement was comparable to<br />

that in treatments containing DNP. This observation suggests that<br />

the drug affected processes in addition to CPR transport at the given<br />

concentration. In fact, when applied at concentrations of 1mmoll –1<br />

or above, probenecid is reported to induce a range of non-specific<br />

effects presumably initiated by the uncoupling of mitochondrial<br />

Haemolymph<br />

Na + /K + -ATPase<br />

K + OA –<br />

ATP<br />

P Na +<br />

i +ADP<br />

Tardigrade<br />

midgut cell<br />

P i +ADP H+ ATP<br />

V-type H + -ATPase<br />

Lumen<br />

Solute carrier<br />

(SLC21/SLCO)<br />

Nucleus<br />

OA –<br />

Basal lamina<br />

Apical junction<br />

complex<br />

Fig.9. Tentative model of the tardigrade midgut cell derived from the<br />

current study on H. crispae. Based on the pharmacological profile of the<br />

tardigrade midgut epithelium, both the Na + /K + -ATPase and the V-type H + -<br />

ATPase are potential candidates for providing an electrochemical driving<br />

force for the transepithelial movement of organic anions. Transport<br />

characteristics and the presence in tardigrade EST libraries suggest that a<br />

member of the SLC21/SLCO transporter family may mediate the<br />

basolateral entry of organic anions in tardigrades. The exact coupling<br />

between electrochemical gradients generated by the pumps and transport<br />

of ions, as well as the nature of the apical exit step, are not known.<br />

oxidative phosphorylation (Masereeuw et al., 2000). It is therefore<br />

possible that the reduction in CPR transport is non-specific when<br />

the drug is applied in concentrations ≥1mmoll –1 . At lower<br />

concentrations of probenecid (


506<br />

K. A. <strong>Halberg</strong> and N. Møbjerg<br />

ID: HDC00004, HDC02687 and HDC03352) (ABCB and ABCC<br />

protein families) are present and expressed in tardigrades – just as<br />

they are in insects (Maddrell et al., 1974; Lanning et al., 1996;<br />

Bresler et al., 1990; Linton and O’Donnell, 2000; Torrie et al., 2004;<br />

Neufeld et al., 2005; Leader and O’Donnell, 2005; O’Donnell and<br />

Leader, 2006; Chahine and O’Donnell, 2009). Comparing the<br />

transport characteristics of the different groups of transporters reveals<br />

that it is unlikely that members of the ABC transporter family (i.e.<br />

P-gps and Mrps) mediate the basolateral entry of CPR in H. crispae<br />

and S. gregaria. This assumption is based on the fact that P-gp<br />

transporters predominantly recognize large cationic species (Russel,<br />

2010), and because Mrp transporters mainly transport large<br />

(>500Da) polyvalent type II organic anions (Wright and Dantzler,<br />

2004; Russel, 2010) – the apical transporter MRP2, however, was<br />

shown to transport PAH in humans (see below). In addition, PAH<br />

does not compete with MTX (an Mrp substrate) in D. melongaster,<br />

suggesting distinct transport systems for these compounds in insects<br />

(Chahine and O’Donnell, 2009). In contrast, both members of the<br />

SLC22 and SLC21/SLCO protein families (Oats and Oatps) have<br />

been reported to transport all the investigated organic anions,<br />

including probenecid, PAH and CPR (Pritchard et al., 1999; Lee<br />

and Kim, 2004; Torrie et al., 2004). Therefore, it seems probable<br />

that an Oat or Oatp homologue mediates the basolateral entry of<br />

CPR in the tardigrade midgut cell as well as the insect MTs (Fig.9).<br />

At present, our data do not allow us to make assumptions on the<br />

nature of the luminal exit. Evidence suggests that MRP2 is involved<br />

in the efflux of organic anions across the brush-border membrane<br />

in the human kidney proximal tubule (Leier et al., 2000), and a<br />

similar situation could exist in tardigrades and insects. Our data<br />

tentatively suggest that the V-type H + -ATPase, and perhaps also<br />

the Na + /K + -ATPase, provide the driving force for the transepithelial<br />

transport of organic anions in both H. crispae and S. gregaria<br />

(Fig.9). It is likely that a large lumen-positive transepithelial<br />

potential, generated by an apical H + -ATPase provides a significant<br />

driving force for the accumulation of anions in the lumen. However,<br />

the exact coupling between electrochemical gradients generated by<br />

the pumps and transport of the ions is not known.<br />

In future studies, it would be of interest to investigate whether<br />

substrates of Mrps (e.g. Texas Red and MTX) also accumulate in<br />

the midgut of tardigrades, and whether transport of these anions<br />

occurs via a separate or a common transporter to CPR. Similarly,<br />

an understanding of the electrophysiological properties of the<br />

midgut epithelium would be highly relevant in our ongoing struggle<br />

to understand the complex biology of these amazing animals.<br />

ACKNOWLEDGEMENTS<br />

We would like to thank Reinhardt M. Kristensen for the use of the Olympus BX50<br />

stereomicroscope, Jette Lyby Michelsen for technical assistance and Dennis K.<br />

Persson and Aslak Jørgensen for help during sampling.<br />

FUNDING<br />

Funding came from the 2008 Faculty of Science, University of Copenhagen Freja-<br />

Programme and from the Carlsberg Foundation.<br />

REFERENCES<br />

Aguinaldo, A. M. A., Turbeville, J. M., Linford, L. S., Rivera, M. C., Garey, J. R., Raff,<br />

R. A. and Lake, J. A. (1997). Evidence for a clade of nematodes, arthropods and other<br />

moulting animals. Nature 387, 489-493.<br />

Al-Fifi, Z. I. A. (2007). Comparative of effect of inhibitors on the ATPase from the<br />

excretory systems of the usherhopper, Poekilocerus bufonius and desert locust,<br />

Schistocerca gregaria. Am. J. Cell Biol. 2, 11-22.<br />

Beyenbach, K. W., Skaer, H. and Dow, J. A. T. (2010). The developmental, molecular,<br />

and transport biology of Malpighian tubules. Annu. Rev. Entomol. 55, 351-374.<br />

Bresler, V. M., Belyaeva, E. A. and Mozhayeva, M. G. (1990). A comparative study on<br />

the system of active transport of organic acids in Malpighian tubules of insects. J. Insect<br />

Physiol. 36, 259-270.<br />

Burckardt, B. C. and Burckardt, G. (2003). Transport of organic anions across the<br />

basolateral membrane of proximal tubule cells. Rev. Physiol. Biochem. Pharmacol. 146,<br />

95-158.<br />

Buss, D. S. and Callaghan, A. (2008). Interaction of pesticides with p-glycoprotein and<br />

other ABC proteins: A survey of the possible importance to insecticides, herbicide and<br />

fungicide resistance. Pestic. Biochem. Physiol. 90, 141-153.<br />

Chahine, S. and OʼDonnell, M. J. (2009). Physiological and molecular characterization of<br />

methotrexate transport by Malpighian tubules of adult Drosophila melanogaster. J. Insect<br />

Physiol. 55, 927-935.<br />

Chahine, S. and OʼDonnell, M. J. (2010). Effects of acute or chronic exposure to dietary<br />

organic anions on secretion of methotrexate and salicylate by Malpighian tubules of<br />

Drosophila melanogaster larvae. Arch. Insect Biochem. Physiol. 73, 128-147.<br />

Dantzler, W. H. (2002). Renal organic anion transport: a comparative and cellular<br />

perspective. Biochim. Biophys. Acta 1566, 169-181.<br />

Dewel, R. A. and Dewel, W. C. (1979). Studies on the tardigrades. J. Morphol. 161, 79-<br />

110.<br />

Dow, J. A. T. and Davies, S. A. (2006). The Malpighian tubule: rapid insights from postgenomic<br />

biology. J. Insect. Physiol. 52, 365-378.<br />

Faria, M., Navarro, A., Luckenbach, T., Piña, B. and Barata, C. (2010). Characterization<br />

of the multixenobiotic resistance (mxr) mechanism in embryos and larvae of the zebra<br />

mussel (Dreissena polymorpha) and studies on its role in tolerance to single and mixture<br />

combinations of toxicants. Aqua. Toxicol. 101, 78-87.<br />

Garret, M. A., Bradley, T. J., Meredith, J. E. and Phillips, J. E. (1988). Ultrastructure of<br />

the Malpighian tubules of Schistocerca gregaria. J. Morphol. 195, 313-325.<br />

George, R. L., Wu, X., Huang, W., Fei, Y. J., Leibach, F. H. and Ganapathy, V. (1999).<br />

Molecular cloning and functional characterization of a polyspecific organic anion<br />

transporter from Caenorhabditis elegans. J. Pharmacol. Exp. Ther. 291, 596-603.<br />

Greven, H. (1982). Homologues or analogues A survey of some structural patterns in<br />

Tardigrada. In Proceedings of the Third International Symposium on the Tardigrada (ed.<br />

D. R. Nelson), pp. 55-76. Johnson City, Tennessee: East Tennessee State University<br />

Press.<br />

Hagenbuch, B. and Gui, C. (2008). Xenobiotic transporters of human organic anion<br />

transporting polypeptides (OATP) family. Xenobiotica 38, 778-801.<br />

<strong>Halberg</strong>, K. A., Persson, D., Møbjerg, N., Wanninger, A. and Kristensen, R. M.<br />

(2009a). Myoanatomy of the marine Tardigrade Halobiotus crispae (Eutardigrada:<br />

Hypsibiidae). J. Morphol. 270, 996-1013.<br />

<strong>Halberg</strong>, K. A., Persson, D., Ramløv, H., Westh, P., Kristensen, R. M. and Møbjerg, N.<br />

(2009b). Cyclomorphosis in Tardigrada: adaption to environmental constraints. J. Exp.<br />

Biol. 212, 2803-2811.<br />

Haugen, B. M., <strong>Halberg</strong>, K. A., Jespersen, Å., Prehn, L. R. and Møbjerg, N. (2010).<br />

Functional characterization of the vertebrate primary ureter: structure and ion transport<br />

mechanisms of the pronephric duct of axolotl larvae (Amphibia). BMC Develop. Biol. 10,<br />

56.<br />

Jönsson, K. I., Rabbow, E., Schill, R. O., Ringdahl, M. H. and Rettberg, P. (2008).<br />

Tardigrades survive exposure to space in low Earth orbit. Curr. Biol. 18, R729-R731.<br />

Kristensen, R. M. (1982). The first record of cyclomorphosis in Tardigrada based on a<br />

new genus and species from Arctic meiobenthos. Z. zool. Syst. Evolut.-forsch. 20, 249-<br />

270.<br />

Lanning, C. L., Fine, R. L., Corcoran, J. J., Ayad, H. M., Rose, R. L. and Abou-Donia,<br />

M. B. (1996). Tobacco budworm P-glycoprotein: biochemical characterization and its<br />

involvement in pesticide resistance. Biochim. Biophys. Acta. 1291, 155-162.<br />

Leader, J. P. and OʼDonnell, M. J. (2005). Transepithelial transport of fluorescent p-<br />

glycoprotein and MRP2 substrates by insect Malpighian tubules: confocal microscopic<br />

analysis of secreted fluid droplets. J. Exp. Biol. 208, 4363-4376.<br />

Lebovitz, R. M., Takeyasu, K. and Fambrough, D. M. (1989). Molecular characterization<br />

and expression of the (Na + + K + )-ATPase alpha-subunit in Drosophila melanogaster.<br />

EMBO J. 8, 193-202.<br />

Lee, W. and Kim, R. B. (2004). Transport and renal drug elimination. Annu. Rev.<br />

Pharmacol. Toxicol. 44, 137-166.<br />

Leier, I., Hummel-Eisenbeiss, J., Cui, Y. and Keppler, D. (2000). ATP-dependent paraaminohippurate<br />

transport by apical multidrug resistance protein MRP2. Kidney Int. 57,<br />

1636-1642.<br />

Linton, S. M. and OʼDonnell, M. J. (2000). Novel aspects of the transport of organic<br />

anions by Malpighian tubules of Drosophila melanogaster. J. Exp. Biol. 203, 3575-3584.<br />

Maddrell, S. H. P., Gardiner, B. O. C., Pilcher, D. E. M. and Reynolds, S. E. (1974).<br />

Active transport by insect Malpighian tubules of acidic dyes and of acylamides. J. Exp.<br />

Biol. 61, 357-377.<br />

Marshall, E. K., Jr and Vickers, J. L. (1923). The mechanism of the elimination of<br />

phenolsulphonephthalein by the kidney - a proof of secretion by the convoluted tubules.<br />

Bull. Johns Hopkins Hosp. 34, 1-6.<br />

Masereeuw, R., van-Pelt, A. P., van-Os, S. H. G., Willems, P. H. G. M., Smits, P. and<br />

Russel, F. G. M. (2000). Probenecid interferes with renal oxidative metabolism: a<br />

potential pitfall in its use as an inhibitor of drug transport. Br. J. Pharmacol. 131, 57-62.<br />

Møbjerg, N. and Dahl, C. (1996). Studies on the morphology and ultrastructure of the<br />

Malpighian tubules of Halobiotus crispae Kristensen 1982 (Eutardigrada). Zool. J. Linn.<br />

Soc. 116, 85-99.<br />

Møbjerg, N., Jørgensen, A., Eibye-Jacobsen, J., <strong>Halberg</strong>, K. A., Persson, D. and<br />

Kristensen, R. M. (2007): New records on cyclomorphosis in the marine eutardigrade<br />

Halobiotus crispae (Eutardigrada: Hypsibiidae). J. Limnol. 66 Suppl.1, 132-140.<br />

Møbjerg, N. M., <strong>Halberg</strong>, K. A., Jørgensen, A., Persson. D., Bjørn, M., Ramløv, H. and<br />

Kristensen, R. M. (2011). Survival in extreme environments – on the current knowledge<br />

of adaptations in tardigrades. Acta Physiologica 202, 409-420.<br />

Mulenga, A., Khumthong, R., Chalaire, K. C., Strey, O. and Teel, P. (2008). Molecular<br />

and biological characterization of the Amblyomma americanum organic anion transporter<br />

polypeptide. J. Exp. Biol. 211, 3401-3408.<br />

Neufeld, D. S. G., Kaufmann, R. and Kurtz, Z. (2005). Specificity of the fluorescein<br />

transport process in Malpighian tubules of the cricket Acheta domesticus. J. Exp. Biol.<br />

208, 2227-2236.<br />

OʼDonnell, M. J. and Leader, J. P. (2006). Changes in fluid secretion rate alter net<br />

transepithelial transport of MRP2 and p-glycoprotein substrates in Malpighian tubules of<br />

Drosophila melanogaster. Arch. Insect Biochem. Physiol. 63, 123-134.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Epithelial transport in tardigrades<br />

507<br />

OʼDonnell, M. J., Ianowski, J. P., Linton, S. T. and Rheault, M. R. (2003). Inorganic and<br />

organic anion transport by insect renal epithelia. Biochim. Biophys. Acta 1618, 194-206.<br />

Pelzer, B., Dastych, H. and Greven, H. (2007). The osmoregulatory/excretory organs of<br />

the glacier-dwelling eutardigrade Hypsibius klebelsbergi Mihelcic, 1959 (Tardigrada).<br />

Mitt. Hamb. Zool. Mus. Inst. 104, 61-72.<br />

Persson, D., <strong>Halberg</strong>, K. A., Jørgensen, A., Ricci, C., Møbjerg, N. and Kristensen, R.<br />

M. (2011). Extreme stress tolerance in tardigrades: surviving space conditions in low<br />

earth orbit. J. Zool. Syst. Evol. Res. 49, 90-97.<br />

Phillips, J. (1981). Comparative physiology of insect renal function. Am. J. Physiol. Regul.<br />

Integr. Comp. Physiol, 241, R241-R257.<br />

Pritchard, J. B., Sweet, D. H., Miller, D. S. and Walden, R. (1999). Mechanism of<br />

organic anion transport across the apical membrane of choroid plexus. J. Biol. Chem.<br />

274, 33382-33387.<br />

Rebecchi, L. and Bertolani, R. (1994). Maturative pattern of ovary and testis in<br />

eutardigrades of fresh-water and terrestrial habitats. Invert. Rep. Dev. 26, 107-117.<br />

Rebecchi, L., Altiero, T., Guidetti, R., Cesari, M., Bertolani, R., Negroni, M. and<br />

Rizzo, A. M. (2008). Tardigrade resistance to space effects: first results of<br />

experiments on the LIFE-TARSE mission on FOTON-M3 (September 2007).<br />

Astrobiology 6, 581-591.<br />

Rost-Roszkowska, M. M., Poprawa, I., Wójtowicz, M. and Kaczmare, L. (2011).<br />

Ultrastructural changes of the midgut epithelium in Isohypsibius granulifer granulifer<br />

Thulin, 1928 (Tardigrada: Eutardigrada) during oogenesis. Protoplasma 248, 405-<br />

414.<br />

Russel, F. G. M. (2010). Transporters: importance in drug absorption, distribution, and<br />

removal. In Enzyme- and Transporter-Based Drug-Drug Interactions (ed. K. S. Pang, A.<br />

Rodrigues and R. M. Peter), pp. 27-49. Heidelberg: Springer.<br />

Schmidt-Rhaesa, A. and Kulessa, J. (2007). Muscular architecture of Milnesium<br />

tardigradum and Hypsibius sp. (Eutardigrada, Tardigrada) with some data on<br />

Ramazzottius oberhaeuseri. Zoomorphology 126, 265-281.<br />

Torrie, L. S., Radford, J. C., Southall, T. D., Kean, L., Dinsmore, A. J., Davies, S. A.<br />

and Dow, J. A. T. (2004). Resolution of the insect ouabain paradox. Proc. Natl. Acad.<br />

Sci. USA. 101, 13689-13693.<br />

Towle, D. W., Paulsen, R. S., Weihrauch, D., Kordylewski, M., Salvador, C., Lignot, J.<br />

H. and Spannings-Pierrrot, C. (2001). Na + +K + -ATPase in gills of the blue crab<br />

Callinectes sapidus: cDNA sequencing and salinity-related expression of -subunit<br />

mRNA and protein. J. Exp. Biol. 204, 4005-4012.<br />

Wang, J., Kean, L., Yang, J., Allan, A. K., Davies, S. A., Herzyk, P. and Dow, J. A. T.<br />

(2004). Function-informed transcriptome analysis of Drosophila renal tubule. Gen. Biol.<br />

5, R69.<br />

Weng, X.-H., Huss, M., Wieczorek, H. and Beyenbach, K. W. (2003). The V-type H + -<br />

ATPase in Malpighian tubules of Aedes aegypti: localization and activity. J. Exp. Biol.<br />

206, 2211-2219.<br />

Wright, S. W. and Dantzler, W. H. (2004). Molecular and cellular physiology of renal<br />

organic cation and anion transport. Physiol. Rev. 84, 987-1049.<br />

Zantke, J., Wolff, C. and Scholtz, G. (2008). Three-dimensional reconstruction of the<br />

central nervous system of Macrobiotus hufelandi (Eutardigrada: Parachela): implications<br />

for the phylogenetic position of Tardigrada. Z. Morphol. 127, 21-36.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Paper III


Acta Physiol 2011, 202, 409–420<br />

REVIEW<br />

Survival in extreme environments – on the current<br />

knowledge of adaptations in tardigrades<br />

N. Møbjerg, 1 K. A. <strong>Halberg</strong>, 1 A. Jørgensen, 2 D. Persson, 1,2 M. Bjørn, 3 H. Ramløv 3 and<br />

R. M. Kristensen 2<br />

1 Department of Biology, University of Copenhagen, Copenhagen, Denmark<br />

2 Natural History Museum of Denmark, University of Copenhagen, Copenhagen, Denmark<br />

3 Department of Science, Systems and Models, University of Roskilde, Roskilde, Denmark<br />

Received 17 October 2010,<br />

revision requested 13 November<br />

2010,<br />

revision received 6 January 2011,<br />

accepted 10 January 2011<br />

Correspondence: N. Møbjerg,<br />

Department of Biology, August<br />

Krogh Building, Universitetsparken<br />

13, DK-2100 Copenhagen Ø,<br />

Denmark.<br />

E-mail: nmobjerg@bio.ku.dk<br />

Abstract<br />

Tardigrades are microscopic animals found worldwide in aquatic as well as<br />

terrestrial ecosystems. They belong to the invertebrate superclade Ecdysozoa,<br />

as do the two major invertebrate model organisms: Caenorhabditis elegans<br />

and Drosophila melanogaster. We present a brief description of the tardigrades<br />

and highlight species that are currently used as models for physiological<br />

and molecular investigations. Tardigrades are uniquely adapted to a<br />

range of environmental extremes. Cryptobiosis, currently referred to as a<br />

reversible ametabolic state induced by e.g. desiccation, is common especially<br />

among limno-terrestrial species. It has been shown that the entry and exit of<br />

cryptobiosis may involve synthesis of bioprotectants in the form of selective<br />

carbohydrates and proteins as well as high levels of antioxidant enzymes and<br />

other free radical scavengers. However, at present a general scheme of<br />

mechanisms explaining this phenomenon is lacking. Importantly, recent<br />

research has shown that tardigrades even in their active states may be extremely<br />

tolerant to environmental stress, handling extreme levels of ionizing<br />

radiation, large fluctuation in external salinity and avoiding freezing by<br />

supercooling to below )20 °C, presumably relying on efficient DNA repair<br />

mechanisms and osmoregulation. This review summarizes the current<br />

knowledge on adaptations found among tardigrades, and presents new data<br />

on tardigrade cell numbers and osmoregulation.<br />

Keywords cell numbers, cryptobiosis, evolution, osmoreglation, supercooling,<br />

tardigrade.<br />

Tardigrades, also known as water bears, are microscopic<br />

metazoans (approx. 0.1–1.2 mm). They were<br />

discovered in the 18th Century with the development of<br />

early microscopes and were first described by the<br />

German zoologist Goeze in 1773, who named them<br />

‘kleiner Wasserbär’ or little water bear because of their<br />

strong resemblance to a little bear (see e.g. Ramazzotti<br />

& Maucci 1983, Nelson 2001, Schill 2010). Shortly<br />

after in 1776, the current name, Tardigrada (from Latin<br />

tardigradus, slow-moving), was given by the Italian<br />

natural scientist Spallanzani (see e.g. Rebecchi et al.<br />

2007). Tardigrades are exceptional among metazoans in<br />

their adaptations to the most extreme environments. As<br />

is also known from selected species of arthropods,<br />

nematodes and rotifers, many species have the ability to<br />

enter cryptobiosis; a state of suspended animation<br />

believed to be ametabolic (Keilin 1959, Clegg 2001).<br />

Corti already noted these adaptations in tardigrades in<br />

1774, when he observed that these animals could be<br />

revived after desiccation (Kinchin 1994). In 1962,<br />

Tardigrada was recognized as a phylum by Ramazzotti<br />

in Il Phylum Tardigrada (Ramazzotti 1962). There are<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x 409


Adaptation to extreme environments in tardigrades Æ<br />

N. Møbjerg et al. Acta Physiol 2011, 202, 409–420<br />

currently approx. 1000 described species (Guidetti &<br />

Bertolani 2005, Degma et al. 2010), but this is likely far<br />

from the real number of tardigrade species, as especially<br />

the marine arthrotardigrades remain relatively unexplored.<br />

Tardigrada belongs to the invertebrate superclade<br />

Ecdysozoa, however; their precise phylogenetic position<br />

is still debated and it is presently not clear whether the<br />

group is more closely related to arthropods and<br />

onychophorans or to the nematodes and nematomorphs<br />

(Fig. 1) (Aguinaldo et al. 1997, Dunn et al. 2008,<br />

Edgecombe 2010). In any case, as noticed in recent<br />

papers, this group has a central position placed in<br />

between the two major invertebrate model organisms –<br />

the nematode Caenorhabditis elegans Maupas, 1900<br />

and the arthropod Drosophila (Sophophora) melanogaster<br />

Meigen, 1830 (Goldstein & Blaxter 2002, Gabriel<br />

& Goldstein 2007).<br />

Little is known about the physiological mechanisms<br />

underlying adaptations to extreme environmental conditions<br />

in tardigrades. Past centuries of tardigrade<br />

research have mainly focused on species descriptions<br />

and morphological investigations related to phylogenetic<br />

analysis. In recent years, however, research in the<br />

field has taken advantage of new molecular tools and an<br />

increasing number of scientists find the group fasci-<br />

Loricifera<br />

Kinorhyncha<br />

Priapulida<br />

ECDYSOZOA<br />

<br />

Nematomorpha<br />

Nematoda*<br />

Tardigrada*<br />

Arthropoda*<br />

Onychophora<br />

PROTOSTOMIA<br />

Brachiopoda<br />

Annelida<br />

LOPHOTROCHOZOA<br />

Mollusca<br />

Rotifera*<br />

Platyhelminthes<br />

DEUTEROSTOMIA<br />

Chordata<br />

Echinodermata<br />

Cnidaria<br />

Figure 1 Evolutionary position of tardigrades in the Animal Kingdom. Phylogeny of the Metazoa (animals) based on<br />

Aguinaldo et al. (1997) and Dunn et al. (2008) showing selected phyla with emphasis on the position of the Tardigrada. The<br />

inferred position of the tardigrades is based on EST sequences from Richtersius coronifer and Hypsibius dujardini. Ecdysozoa<br />

includes all molting animals and is one of the two protostome superclades. The marked phyla have cryptobiotic species.<br />

Porifera<br />

410<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x


Acta Physiol 2011, 202, 409–420<br />

N. Møbjerg et al. Æ<br />

Adaptation to extreme environments in tardigrades<br />

nating. In particular, the mechanisms underlying the<br />

ability to enter cryptobiosis have attracted considerable<br />

scientific interest. This interest is no doubt associated<br />

with the suspected translational output related to a<br />

detailed understanding of the complex stress physiology<br />

of tardigrades, i.e. in connection with cryopreservation<br />

and dehydration of biological material. Noticeably, the<br />

phenomenon cryptobiosis touches upon our conception<br />

of life and death; one of the largest enigmas being how<br />

metabolism is restarted after years of suspension. This<br />

review puts focus on physiological and molecular<br />

adaptations to extreme conditions found among tardigrades.<br />

We additionally present a brief description of<br />

the phylum and highlight species that are currently used<br />

as models for these investigations.<br />

Phylum Tardigrada<br />

Tardigrade phylogeny and evolution<br />

About 35 extant animal groups have body plans and<br />

genes that are distinct enough to warrant elevation to<br />

phylum status (Nielsen 2001). The tardigrades, comprising<br />

the phylum Tardigrada, are one of these groups<br />

(Fig. 1). Tardigrades are microscopic invertebrates with<br />

a well developed organization including brain and<br />

sensory organs, muscles, a complex feeding apparatus<br />

and alimentary tract, reproductive and osmoregulatory<br />

organs (see e.g. Rebecchi & Bertolani 1994, Dewel &<br />

Dewel 1996, Møbjerg & Dahl 1996, Eibye-Jacobsen<br />

1997, Jørgensen et al. 1999, Greven 2007, <strong>Halberg</strong><br />

et al. 2009a). They are found worldwide in aquatic as<br />

well as terrestrial environments, but depend on free<br />

water to be in their active, reproducing state. It has been<br />

suggested that tardigrades, like e.g. nematodes, have<br />

eutely, but detailed studies on the subject are still<br />

lacking. Cell counts based on nuclear staining with<br />

DAPI (4¢,6-diamidino-2-phenylindole) in four active<br />

stage adults of the marine eutardigrade Halobiotus<br />

crispae Kristensen, 1982 revealed a total cell number of<br />

around 1060 cells, when excluding gametes (Fig. 2).<br />

This number could represent a slight underestimate of<br />

the total somatic cell number as body cavity cells (also<br />

known as storage cells) may have escaped two of the<br />

specimens, which were punctured prior to the staining<br />

(Fig. 2). We did not observe cell divisions (mitosis)<br />

during the counts. Mitosis has, however, previously<br />

been reported in post-embryonic eutardigrades (Bertolani<br />

1970a,b, 1982).<br />

There are two main evolutionary lines within the<br />

tardigrades, represented by the classes Eutardigrada and<br />

Heterotardigrada (Fig. 3) (see e.g. Jørgensen & Kristensen<br />

2004). The validity of a third class, Mesotardigrada,<br />

is currently uncertain. Mesotardigrada only<br />

contains a single species, Thermozodium esakii Rahm,<br />

1937 originally found in a hot spring in Japan. The type<br />

specimens of T. esakii no longer exist and the type<br />

locality was apparently destroyed in an earthquake<br />

A<br />

P<br />

br<br />

pb<br />

gI<br />

c.gl.<br />

Leg 1<br />

eo<br />

gII<br />

c.gl.<br />

Leg 2<br />

mg<br />

gIII<br />

c.gl.<br />

Leg 3<br />

go<br />

gIV<br />

c.gl.<br />

Leg 4<br />

Figure 2 Cell numbers in Halobiotus crispae. 3-D reconstruction of cell arrangement in the eutardigrade H. crispae based on<br />

confocal laser scanning microscopy of a DAPI stained specimen. In order to obtain an estimate of somatic cell numbers in this<br />

species, specimens were relaxed in CO 2 -enriched water, fixed in 4% paraformaldehyde and ultrasonicated. Two of the four<br />

specimens were additionally delicately punctured with a fine needle. The specimens were subsequently incubated with DAPI and<br />

thoroughly rinsed before mounting. Image acquisition was performed using a Leica DM RXE 6 TL microscope equipped with a<br />

Leica TCS SP2 AOBS confocal laser scanning unit. Cell counts and image processing were performed using the software program<br />

Imaris (Bitplane, Zurich, Switzerland). The total number of somatic cells in H. crispae was estimated at approx. 1060, based on<br />

stainings of four male specimens (mean SD: 1058 53). This number could represent a slight underestimate of the total somatic<br />

cell number as so-called body cavity cells may have escaped the two specimens that were punctured prior to staining (cell counts for<br />

the punctured specimens: 998 and 1088; counts for the non-punctured specimens: 1036 and 1112). The largest number of cells is<br />

clearly present in the anterior part of the animal containing the brain and buccopharyngeal apparatus. A, anterior; P, posterior; br,<br />

brain; c.gl., claw gland; eo, esophagus; gI–IV, ventral ganglia I–IV; mg, midgut; go, gonad; pb, pharyngeal bulb. Scale bar: 50 lm.<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x 411


Adaptation to extreme environments in tardigrades Æ<br />

N. Møbjerg et al. Acta Physiol 2011, 202, 409–420<br />

Echiniscoides<br />

Echiniscoidae<br />

Echiniscus<br />

Echiniscidae<br />

Milnesium*<br />

Milnesiidae<br />

Halobiotus<br />

Isohypsibioidea<br />

Richtersius<br />

Paramacrobiotus<br />

Macrobiotoidea<br />

Hypsibius*<br />

Ramazzottius*<br />

Hypsibioidea<br />

Arthrotardigrada<br />

marine<br />

non-cryptobionts<br />

Echiniscoidea<br />

limno-terrestrial<br />

Intertidal<br />

Cryptobionts<br />

Apochela<br />

limno-terrestrial<br />

Cryptobionts<br />

Parachela<br />

limno-terrestrial<br />

cryptobionts<br />

non-cryptobionts<br />

Heterotardigrada<br />

Mesotardigrada ()<br />

Eutardigrada<br />

Tardigrada<br />

*Genome projects<br />

Figure 3 Tardigrade phylogeny. Phylogeny<br />

of tardigrades showing major clades<br />

and position of model/discussed species.<br />

The phylogeny is based on Sands et al.<br />

(2008).<br />

(Nelson 2002). However, a thorough re-sampling for<br />

this species has to our knowledge not been performed.<br />

Tardigrades most likely evolved within the marine<br />

environment, and marine species are especially numerous<br />

within the heterotardigrade order Arthrotardigrada<br />

(Renaud-Mornant 1982, Maas & Waloszek 2001,<br />

Jørgensen et al. 2010). Arthrotardigrades are present<br />

in all oceans from intertidal zones to abyssal depths,<br />

inhabiting different sediment types. In addition, marine<br />

species are found within the other main heterotardigrade<br />

order, Echiniscoidea, represented by the intertidal<br />

Echiniscoides sigismundi (M. Schultze, 1865). This<br />

species may very well be the toughest creature on<br />

Earth, having to endure periods of desiccation and low<br />

oxygen tension as well as large perturbations in salinity<br />

and freezing (Kristensen & Hallas 1980). Nevertheless,<br />

the exact range of this tardigrade’s tolerances remains<br />

to be investigated. It may be hypothesized that an<br />

Echiniscoides-like tardigrade invaded the freshwater/<br />

terrestrial environment and gave rise to the almost<br />

exclusively limno-terrestrial eutardigrades (Kinchin<br />

1994). This is however currently not supported by<br />

molecular data (Sands et al. 2008, Jørgensen et al.<br />

2010; Jørgensen et al. 2011).<br />

The eutardigrades are divided into two orders;<br />

Apochela and Parachela. Two genera within the latter<br />

order, Ramajendas represented by Ramajendas renaudi<br />

(Ramazzotti 1972) in the Southern Hemisphere and<br />

Halobiotus in the Northern Hemisphere, have secondarily<br />

invaded the marine environment (Ramazzotti<br />

1972, Kristensen 1982, Møbjerg et al. 2007). Although<br />

cryptobiosis is common in most eutardigrades, our<br />

recent findings in H. crispae suggest that among these<br />

secondary marine species, adaptations are present that<br />

are quite extraordinary (<strong>Halberg</strong> et al. 2009b). The<br />

tardigrades stay active while experiencing large fluctuations<br />

in abiotic factors, fluctuations that in other<br />

tardigrades would induce cryptobiosis.<br />

Tardigrade genomes<br />

There is a huge variation in the genome size of<br />

tardigrades ranging from about 75–100 Mb in Hypsibius<br />

and Ramazzottius to 800 Mb in Bertolanius<br />

(Gregory 2010, C-values converted from picograms to<br />

base pairs using the conversion 1 pg = 978 Mb according<br />

to Dolezel et al. (2003), Bertolanius was previously<br />

named Amphibolus). For comparison, the genome sizes<br />

of Caenorhabditis elegans and Drosophila melanogaster<br />

are about 100 Mb and 175 Mb respectively (Gregory<br />

2010). The general diploid chromosome numbers of<br />

the eutardigrades are 10–12 and 14 for the heterotardigrade<br />

Echiniscus (Bertolani 1982). Polyploidy with<br />

up to 24 chromosomes is common in eutardigrades<br />

(Bertolani 1982). Three major sequencing projects are<br />

currently ongoing within Tardigrada. However, all<br />

three projects are investigating eutardigrade species;<br />

no data are presently available for the other main<br />

tardigrade group – the heterotardigrades. The international<br />

collaborative Ecdysozoan Sequencing Project is<br />

assembling the genome of Hypsibius dujardini (Doyère,<br />

1840) as part of an investigation into the ancestral<br />

genome of the Ecdysozoa lineage. Prior to this project,<br />

the Edinburgh based TardiBASE project, generated<br />

more than 5000 EST sequences for H. dujardini<br />

(GenBank 2010). The German based FUNCRYPTA<br />

project was focused on investigating cryptobiosis in<br />

Milnesium tardigradum Doyère, 1840 through studies<br />

of gene and protein expression (Förster et al. 2009,<br />

Mali et al. 2010, Schokraie et al. 2010). The project<br />

had in 2010 generated approx. 7000 quality EST<br />

sequences and aimed at advancing the basic understanding<br />

of protein expression in tardigrades through transcriptomic<br />

and proteomic studies (Mali et al. 2010). The<br />

Japanese based Kumamushi Genome Project is assembling<br />

the genome of Ramazzottius varieornatus Bertolani<br />

and Kinchin, 1993 and has preliminarily predicted<br />

412<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x


Acta Physiol 2011, 202, 409–420<br />

N. Møbjerg et al. Æ<br />

Adaptation to extreme environments in tardigrades<br />

about 20 000 candidate genes in this species (Horikawa<br />

et al. 2008, Katayama et al. 2009). The great challenge<br />

in the years to come will be to correlate the description of<br />

tardigrade genes with studies on the function of these<br />

genes as nicely illustrated in a recent study on a<br />

purinergic P2X receptor from H. dujardini (Bavan et al.<br />

2009).<br />

Adaptation to extreme environments<br />

Tardigrades are extraordinary in their tolerance to<br />

extremes, including limno-terrestrial habitats that frequently<br />

dry out, habitats that freeze and habitats that<br />

experience large fluctuations in e.g. osmotic pressure<br />

and oxygen tension. Cryptobiosis, referred to as a<br />

reversible ametabolic state induced by unfavorable<br />

environmental conditions, is a common adaptation<br />

especially among limno-terrestrial tardigrade species<br />

(see e.g. Wright 2001). Four cryptobiosis inducing<br />

physical extremes are traditionally recognized: dehydration<br />

(anhydrobiosis), extremely low temperatures<br />

(cryobiosis), lack of oxygen (anoxybiosis) and high salt<br />

concentration (osmobiosis) (Keilin 1959), with desiccation<br />

induced anhydrobiosis and freezing induced cryobiosis<br />

being the most extensively studied states.<br />

Anhydrobiosis and cryobiosis are not equivalent phenomena<br />

and likely involve different mechanisms for<br />

protection of cells and tissues (Crowe et al. 1990,<br />

1992). Little is known of cryobiosis in tardigrades<br />

(Westh et al. 1991, Ramløv & Westh 1992, Westh &<br />

Kristensen 1992, <strong>Halberg</strong> et al. 2009b, Hengherr et al.<br />

2009, 2010), whereas a great deal of attention has been<br />

paid to anhydrobiosis (Wright et al. 1992, Wright<br />

2001, Rebecchi et al. 2007, Schill 2010). The anhydrobiotic<br />

state is characterized by the formation of a socalled<br />

tun with withdrawn legs and a longitudinally<br />

contracted body (see e.g. Bertolani et al. 2004). Tun<br />

formation is also seen in bdelloid rotifers, whereas<br />

desiccation tolerant nematodes coil into a tight spiral.<br />

Obviously, the ability to pack internal organs during<br />

tun formation is an important adaptation to desiccation.<br />

Importantly, tun formation is an active, regulated<br />

event and not merely an effect of water removal (Crowe<br />

1972). Along this line, our unpublished data show that<br />

when we expose the active state of the cryptobiotic<br />

tardigrade Richtersius coronifer (Richters, 1903) to<br />

water containing high levels of chemical substances, the<br />

tardigrades will respond by contracting their bodies<br />

initiating tun formation thus undergoing chemobiosis–a<br />

cryptobiotic response to environmental toxins. The<br />

formation of the tun is a critical and necessary event<br />

for tardigrades entering anhydrobiosis. Much more<br />

work needs to be done in order to understand the<br />

processes undertaken during transformation to this<br />

ametabolic state, from the molecular to whole organism<br />

level, anatomically as well as physiologically. In the tun<br />

state tardigrades may not only stand long periods of<br />

desiccation and exposure to toxic chemicals but also<br />

very low subzero temperatures, vacuum, high pressure,<br />

radiation, extreme pH, anoxia and to some extent high<br />

temperature (see e.g. Wright 2001, Rebecchi et al.<br />

2007). It has been suggested and to some degree shown<br />

that adaptations to these extremes may involve synthesis<br />

of bioprotectants in the form of selective carbohydrates<br />

and proteins, high levels of antioxidant enzymes<br />

and other free radical scavengers, biological membranes<br />

containing specific phospholipids as well as powerful<br />

DNA repair mechanisms (Westh & Ramløv 1991, Schill<br />

et al. 2004, Jönsson et al. 2005, Rizzo et al. 2010).<br />

In addition to cryptobiotic tardigrade species, we also<br />

find species that form cysts and enter diapause (see e.g.<br />

Møbjerg et al. 2007, Guidetti et al. 2008). Importantly,<br />

it has been shown that tardigrades even in their active<br />

states may be extremely tolerant to environmental stress<br />

(May et al. 1964, Jönsson et al. 2005, Horikawa et al.<br />

2006, <strong>Halberg</strong> et al. 2009b). Virtually nothing is known<br />

about the normal physiology of tardigrades, and<br />

answers to how they tolerate these extremes may<br />

actually be found here. We have recently shown that<br />

the littoral eutardigrade Halobiotus crispae handles<br />

large fluctuation in external salinity and avoids freezing<br />

by supercooling to around )20 °C in its active stage<br />

(<strong>Halberg</strong> et al. 2009b). This species is characterized by<br />

seasonal cyclic changes in morphology and physiology<br />

known as cyclomorphosis, one of the cyclomorphic<br />

stages being freeze tolerant (Kristensen 1982, Møbjerg<br />

et al. 2007, <strong>Halberg</strong> et al. 2009b). The morphological<br />

changes occurring during cyclomorphosis in H. crispae<br />

in some respects resemble the formation of dauer larvae<br />

in C. elegans (see e.g. Cassada & Russell 1975, Burnell<br />

et al. 2005).<br />

Cryptobiosis in tardigrades<br />

Many experiments on tardigrade cryptobiosis have been<br />

performed on the eutardigrade Richtersius coronifer.<br />

R. coronifer, also known as the giant yellow water bear,<br />

has a body length of up to 1 mm (Fig. 4a–c). Both males<br />

and females are present in some populations, but, as<br />

commonly found among eutardigrades, several populations<br />

reproduce by parthenogenesis. R. coronifer lives in<br />

moss in alpine and arctic environments and is furthermore<br />

numerous in moss on carbonated bedrock in dry<br />

Swedish lowland areas known as Alvar (see e.g. Westh<br />

& Kristensen 1992, Jönsson et al. 2001). R. coronifer is<br />

a true cryptobiont, tolerating extreme desiccation as<br />

imposed by e.g. space vacuum conditions (Jönsson et al.<br />

2008, Persson et al. 2011) and it additionally survives<br />

exposures to very low temperatures as encountered by<br />

transfers into liquid nitrogen (approx. )196 °C) in the<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x 413


Adaptation to extreme environments in tardigrades Æ<br />

N. Møbjerg et al. Acta Physiol 2011, 202, 409–420<br />

(a) (b) (c)<br />

st<br />

pb<br />

eo<br />

A<br />

P<br />

A<br />

P<br />

A<br />

P<br />

mg<br />

st<br />

pb mg<br />

*<br />

*<br />

*<br />

(d) (e) (f)<br />

st<br />

st<br />

st<br />

pb<br />

eo<br />

pb<br />

eo<br />

pb<br />

mg<br />

mg<br />

mg<br />

go<br />

* * * *<br />

*<br />

* * *<br />

*<br />

Figure 4 Cryptobiotic and non-cryptobiotic survival in extreme environments. (a–c): Microscopy of Richtersius coronifer from<br />

Stora Alvar on the Swedish island Öland in the Baltic Sea. Richtersius coronifer has been a model tardigrade for investigations on<br />

cryptobiotic survival. (a) Active state; (b) light microscopy and (c) scanning electron microscopy of the cryptobiotic tun state<br />

induced by desiccation. In this state R. coronifer e.g. tolerates complete desiccation as experienced by vacuum condition and<br />

freezing in liquid nitrogen. (d–f): Microscopy of Halobiotus crispae from the Danish population at Vellerup Vig, Denmark. This<br />

tardigrade is one of a few species of eutardigrades that have secondarily invaded the marine environment. Halobiotus crispae in the<br />

active stage (d) uniquely adapted to cope with profound changes in ambient salinity occurring in tidal and subtidal habitats. Upon<br />

transfers to dilute salt water solutions, active stage H. crispae swell (e) and subsequently regulate their body volume to near control<br />

conditions. Halobiotus crispae is characterized by appearing in different cyclomorphic stages. The pseudosimplex 1 stage (f) is freeze<br />

tolerant. eo, esophagus; mg, midgut; pb, pharyngeal bulb; st, stylet. Asterisks mark the position of three Malpighian tubules<br />

presumably involved in osmoregulation. Scale bars: 50 lm.<br />

tun as well as active hydrated state, however,with the<br />

tuns tolerating considerable longer time of exposure<br />

(Ramløv & Westh 1992, Persson et al. 2011). The<br />

anhydrobiotic state survives temperatures of up to<br />

approx. 70 °C for 1 h, but survival rapidly decreases<br />

when the temperature exceeds 70 °C, and no specimens<br />

survives exposure to 100 °C (Ramløv & Westh 2001).<br />

The increase in life span offered by anhydrobiosis in this<br />

species seems to be restricted to approx. 5 years (Westh<br />

& Kristensen 1992). This is less than the cryptobiotic<br />

life expansion observed for the heterotardigrade Echiniscus<br />

testudo (Doyère, 1840), for which we have<br />

revived specimens from moss cushions dried for approx.<br />

20 years (Jørgensen et al. 2007). What then sets the<br />

limits for cryptobiotic survival Probably accumulated<br />

damage to DNA and other molecules as well as to<br />

tissues and organs obtained during the ametabolic state.<br />

This includes damage obtained through oxidative processes<br />

as well as predation, bacterial and fungal<br />

infections. A recent investigation in Paramacrobiotus<br />

richtersi (Murray, 1911) has shown that regulation of<br />

antioxidant metabolism likely plays an important role<br />

in defense against the potential oxidative damage<br />

associated with dehydration (Rizzo et al. 2010). It has<br />

long been known that there is a positive correlation<br />

between the time spent in the anhydrobiotic state and<br />

the time required for recovering active life after rehydration<br />

(Crowe & Higgins 1967). Recent investigations<br />

indicate that the longer the time spent in anhydrobiosis,<br />

the more damage is inflicted to DNA (Neumann et al.<br />

2009, Rebecchi et al. 2009b), which would explain the<br />

prolonged recovery time, indicating that considerable<br />

repair of DNA (and other molecules) occurs in the postanhydrobiotic<br />

state.<br />

Trehalose accumulation. Current knowledge suggests<br />

that specific carbohydrates and proteins are important<br />

in protecting the cells from damage encountered during<br />

414<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x


Acta Physiol 2011, 202, 409–420<br />

N. Møbjerg et al. Æ<br />

Adaptation to extreme environments in tardigrades<br />

entry and exit of cryptobiosis. Bioprotectants may<br />

interact directly with macromolecular structures such<br />

as membranes, DNA and proteins, or they may act as<br />

osmolytes during osmotic or dehydration stress. As has<br />

been shown for other cryptobiotic invertebrates, noticeably<br />

in the cysts of the brine shrimp Artemia franciscana<br />

Kellogg, 1906, some, but not all, cryptobiotic<br />

tardigrades accumulate the dissacharide trehalose during<br />

desiccation (Clegg 1965, Westh & Ramløv 1991,<br />

Hengherr et al. 2008, Jönsson & Persson 2010). Trehalose<br />

has been proposed to act as a molecular stabilizer<br />

replacing water and to further stabilize cellular structure<br />

through the formation of amorphous glasses, a<br />

process known as vitrification (Clegg 2001). The highest<br />

trehalose concentrations measured in anhydrobiotic<br />

tardigrades ranges from 2.3% d.w. in R. coronifer<br />

and Macrobiotus krynauwi Dastych and Harris, 1995<br />

to 2.9% d.w. in Macrobiotus islandicus Richters, 1904<br />

(Westh & Ramløv 1991, Jönsson & Persson 2010).<br />

These trehalose values are, however, relatively low<br />

when compared with the above mentioned crustacean,<br />

which accumulates the disaccharide in concentrations of<br />

around 15% d.w. Moreover, trehalose levels are barely<br />

measurable and do not increase during dehydration in<br />

the cryptobiotic Milnesium tardigradum (see Hengherr<br />

et al. 2008, Jönsson & Persson 2010). The latter<br />

investigations, analysing trehalose contents in eutardigrades<br />

as well as heterotardigrades, show that<br />

trehalose accumulation does not represent a universal<br />

protective mechanism enabling tardigrades to undergo<br />

cryptobiosis.<br />

Expression of heat-shock proteins. Cryptobiosis has<br />

been suggested to rely on the synthesis of molecular<br />

chaperones such as heat-shock-proteins, which may<br />

assist folding of newly synthesized proteins, control<br />

their final intracellular location, as well as protect them<br />

from stress-associated denaturation and aid in renaturation<br />

(Clegg 2001). During entrance into anhydrobiosis<br />

Ramløv & Westh (2001) described the upregulation of<br />

a protein with a molecular weight of approx. 71 kDa in<br />

R. coronifer and proposed that this protein might<br />

belong to the heat-shock-protein (Hsp70) family. Schill<br />

et al. (2004) subsequently investigated RNA expression<br />

patterns of three Hsp70 isoforms in active and anhydrobiotic<br />

states of M. tardigradum as well as during<br />

entrance into and exit out of anhydrobiosis. Only one<br />

of these isoforms (isoform 2) was significantly induced<br />

by the transition from the active to cryptobiotic state<br />

and interestingly, it showed a considerable increase in<br />

expression during the post-cryptobiotic phase, while<br />

the other isoforms were down regulated during cryptobiosis<br />

as well as in transitional states. Jönsson &<br />

Schill (2007) used an immuno-westernblot method to<br />

quantify the induction of Hsp70 in R. coronifer in<br />

response to desiccation, ionizing radiation and heating.<br />

They reported elevated levels of Hsp70 following both<br />

heat and radiation treatment as well as in tardigrades<br />

rehydrated after a period of desiccation. Noticeably,<br />

the authors found that tardigrades in the desiccated<br />

state had reduced Hsp70 levels as compared to the nontreated<br />

control group and accordingly suggested that<br />

Hsp70 may be involved in the repair processes after<br />

desiccation rather than acting as a biochemical stabilizer<br />

in the dry state. Based on the M. tardigradum EST<br />

library, several additional heat-shock proteins have<br />

been identified, including two a-crystalline heat-shock<br />

proteins, and the relative abundance of the transcripts<br />

coding for these stress proteins have been investigated<br />

during phases of dehydration and rehydration (Reuner<br />

et al. 2010). The results obtained suggested a limited<br />

role for heat-shock proteins in the desiccation tolerance<br />

of M. tardigradum. The authors found a variable<br />

pattern of expression with most of the candidate genes<br />

being down regulated, and only one of the genes (Mthsp90),<br />

being significantly upregulated in the dehydrated<br />

state (Reuner et al. 2010). Comparable studies<br />

in Paramacrobiotus richtersi did not show evidence for<br />

an increased expression of either Hsp70 or Hsp90<br />

between hydrated and dehydrated animals (Rizzo et al.<br />

2010). Additionally, specimens of P. richtersi sent into<br />

space (Foton-M3 mission) revealed no significant<br />

change as compared to ground controls in the expression<br />

of these heat-shock proteins (Rebecchi et al.<br />

2009a). These contrasting results indicate that Hsp<br />

expression is species specific. A general role in the<br />

cryptobiotic survival of tardigrades can at present not<br />

be attributed to these stress proteins. Nevertheless, a<br />

synergistic effect between HSPs and other bioprotectants<br />

such as trehalose or LEA (late-embryogenesis<br />

abundant) proteins might exist (see e.g. Goyal et al.<br />

2005). Notably, LEA proteins have been described in<br />

many other desiccation tolerant organisms from plants<br />

to arthropods and might analogously be functionally<br />

important in tardigrades (see e.g. Schill 2010, Warner<br />

et al. 2010). Indeed, a putative LEA protein has<br />

recently been detected in M. tardigradum (Schokraie<br />

et al. 2010). Consequently, future studies on these and<br />

other bioprotectants might offer additional clues in our<br />

search to understand the phenomenon cryptobiosis.<br />

Obviously, cryptobiosis is far from understood and<br />

much more research is needed in order to understand<br />

the mechanisms underlying this state. Interestingly, a<br />

preliminary study from the Kumamushi Genome<br />

Project on R. varieornatus suggests that comparative<br />

metabolome profiling of active and anhydrobiotic<br />

states may provide a novel insight into anhydrobiosis<br />

(Arakawa et al. 2009). Additionally, clues to cryptobiotic<br />

survival may be found in the extreme tolerance<br />

seen in active state tardigrades.<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x 415


Adaptation to extreme environments in tardigrades Æ<br />

N. Møbjerg et al. Acta Physiol 2011, 202, 409–420<br />

Non-cryptobiotic survival in tardigrades<br />

While tardigrades are well known for their abilities to<br />

cope with extreme environmental conditions by entering<br />

cryptobiosis, little focus has been on their ability to<br />

sustain metabolism and remain active during fluctuating<br />

external circumstances. In this context it is interesting to<br />

note that tardigrades are genuine aquatic animals – they<br />

are dependent on free water to be in their active feeding<br />

and reproducing states. Truly terrestrial conditions are<br />

only survived following entry into the cryptobiotic<br />

state. Furthermore, limno-terrestrial tardigrades that<br />

have the ability to enter the tun state may do so in<br />

response to environmental challenges (e.g. exposure to<br />

chemical substances) that in marine species would not<br />

force the animals into cryptobiosis. In the following we<br />

discuss what is currently known about the physiology<br />

and stress tolerance of tardigrades in their active states.<br />

Radiation tolerance. Investigations on radiation tolerance<br />

in Richtersius coronifer have revealed that exposure<br />

to c-radiation at doses up to 1 kGy does not affect<br />

survival of desiccated nor hydrated animals, with<br />

hydrated animals tolerating doses of up to 5 kGy (Jönsson<br />

et al. 2005). Horikawa and co-workers revealed that<br />

both hydrated and desiccated Milnesium tardigradum<br />

survive doses of c-radiation of more than 5 kGy, and up to<br />

8 kGy of heavy ion radiation (Horikawa et al. 2006).<br />

These and previous studies indicate that hydrated animals<br />

are just as good or even better at tolerating radiation. This<br />

indicates that radiation tolerance is not due to biochemical<br />

protectants associated with the cryptobiotic state, but<br />

suggests that tardigrades rely on efficient and yet unidentified<br />

mechanisms of DNA repair (Jönsson et al. 2005,<br />

Horikawa et al. 2006). Several recent studies involved<br />

with the BIOPAN 6/Foton-M3 mission in 2007, funded<br />

by the European Space Agency, have investigated the<br />

impact on survival of tardigrades exposed to space<br />

conditions (Jönsson et al. 2008, Rebecchi et al. 2009a,<br />

Persson et al. 2011). During this mission, cryptobiotic<br />

tardigrades (as well as nematodes and rotifers) were sent<br />

into low earth orbit and exposed to space vacuum and<br />

cosmic radiation. The three studies revealed discrepancies<br />

in survival rates likely reflecting differences between<br />

tardigrade species and experimental setups, however;<br />

they unanimously conclude that tardigrades can survive<br />

the rigors of space, and that M. tardigradum is likely the<br />

most resistant species with embryos as well as adults<br />

tolerating space conditions (see discussion in Persson<br />

et al. 2011). Detailed knowledge of the life history of this<br />

predatory tardigrade is available as a result of a comprehensive<br />

study by Suzuki (2003).<br />

Osmoregulation. Experiments on the marine eutardigrade<br />

Halobiotus crispae have revealed an extraordinary<br />

tolerance to perturbations in body volume and osmotic<br />

pressure (<strong>Halberg</strong> et al. 2009b). The tidal habitat in<br />

which this tardigrade lives is characterized by large<br />

fluctuations in abiotic factors; most noticeably, alterations<br />

are seen in salinity and temperature (Møbjerg et al.<br />

2007). One way of coping with these extremes would be<br />

to enter cryptobiosis. However, H. crispae is not a<br />

cryptobiont. In the active stage the tardigrade handles<br />

extremes by expending energy on active regulatory<br />

mechanisms. In addition, the so-called pseudosimplex 1<br />

(P1) stage of this animal is freeze tolerant (<strong>Halberg</strong> et al.<br />

2009b); it is not, however, a cryobiont as the P1 stage does<br />

not tolerate gradual freezing to )80 °C. Cryobiosis is<br />

defined by the apparent absence of a lower lethal<br />

temperature (Wright 2001). The P1 stage is distinctly<br />

characterized by a double cuticle and closed mouth and<br />

cloaca (Kristensen 1982, Møbjerg et al. 2007).<br />

Halobiotus crispae kept at a salinity of 20 ppt have an<br />

extensive ability to supercool (avoid freezing) down to<br />

around )20 °C, enabling active stage animals to withstand<br />

subzero temperatures without freezing (<strong>Halberg</strong><br />

et al. 2009b). Similar supercooling points have been<br />

reported from limno-terrestrial eutardigrades (Macrobiotus,<br />

Paramacrobiotus and Milnesium), whereas heterotardigades<br />

(Echiniscus) had slightly higher points<br />

(Hengherr et al. 2009). Much higher supercooling<br />

points of )6.7 °C and )7.4 °C were originally observed<br />

in respectively Richtersius coronifer and Bertolanius<br />

nebulosus (Westh & Kristensen 1992). These high<br />

supercooling points are likely the result of the presence<br />

of ice-nucleating agents initiating the freezing process<br />

(Westh et al. 1991). Figure 5 shows thermograms of H.<br />

crispae kept in seawater with a salinity of 20 ppt<br />

(Fig. 5a) and distilled water (Fig. 5b) prior to differential<br />

scanning calorimetry (DSC). Water content in the<br />

animals kept at 20 ppt and in distilled water was<br />

respectively 73% and 81%. The latter group of tardigrades<br />

had a higher supercooling point as compared to<br />

the group kept in 20 ppt saltwater, illustrating the<br />

expected correlation between water content and crystallization<br />

temperature in species without ice-nucleating<br />

agents (see e.g. Hengherr et al. 2009). Thus, the extent<br />

of supercooling and thereby the ability to avoid freezing,<br />

is coupled to osmoregulation.<br />

Halobiotus crispae has a large capacity to tolerate<br />

perturbations in ambient salinity making it an ideal<br />

model for the study of osmoregulation and volume<br />

regulation in tardigrades. Experiments on this species<br />

revealed tolerance to a wide range of salinities, with<br />

specimens of the Greenlandic population remaining<br />

active in solutions ranging from distilled water to<br />

saltwater with osmolalities up to 2000 mOsm kg )1<br />

(<strong>Halberg</strong> et al. 2009b). During experiments with transfers<br />

to strong hypotonic solutions, active stage H.<br />

crispae swell with up to 60% before regulating back to<br />

416<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x


Acta Physiol 2011, 202, 409–420<br />

N. Møbjerg et al. Æ<br />

Adaptation to extreme environments in tardigrades<br />

(a) 15<br />

(a) 1400<br />

Differential heat flow (mW)<br />

Differential heat flow (mW)<br />

7.5<br />

0<br />

(b) 15<br />

7.5<br />

0<br />

Exotherm Exotherm<br />

Cooling<br />

Heating<br />

near control values (Fig. 4e). Similarly, specimens<br />

transferred into hypertonic solutions shrink and thereafter<br />

respond by regulating body volume. In a series of<br />

experiments with transfers into saltwater solutions with<br />

a salinity between 2 ppt (62 mOsm kg )1 ) and 40 ppt<br />

(1245 mOsm kg )1 ) the active stage tardigrades hyperregulated<br />

at all times (Fig. 6a) (<strong>Halberg</strong> et al. 2009b).<br />

Hyperregulation is likely a general feature of at least the<br />

eutardigrades, as our data on R. coronifer reveal that<br />

this species also keeps its body fluids hyperosmotic as<br />

compared to the surroundings (Fig. 6b). Our data<br />

indicate that this limno-terrestrial cryptobiont is less<br />

tolerant of high salinity solutions than the littoral H.<br />

crispae. When exposed to water with increasing salt<br />

Endotherm<br />

20 ppt.<br />

–40 –35 –30 –25 –20 –15 –10 –5 0 5<br />

Cooling<br />

Heating<br />

Endotherm<br />

0 ppt.<br />

–40 –35 –30 –25 –20 –15 –10 –5 0 5<br />

Temperature (°C)<br />

Figure 5 Supercooling in Halobiotus crispae. (a) Supercooling<br />

in H. crispae kept in seawater with a salinity of 20 ppt.<br />

Thermogram modified from <strong>Halberg</strong> et al. (2009b). (b) Halobiotus<br />

crispae kept in distilled water. Cooling and subsequent<br />

reheating of a sample containing 46 specimens of H. crispae<br />

from Vellerup Vig, Denmark kept in distilled water for 45 min<br />

prior to the transfer into pans for differential scanning calorimetry<br />

(DSC). Freezing ()12.2 °C) and melting ()1.2 °C)<br />

temperatures were estimated as onsets of peaks of respectively<br />

exothermic and endothermic events using DSC7 software<br />

(cooling rate 2 °C min )1 ) (DSC7; Perkin Elmer Inc., Wellesley,<br />

MA, USA).<br />

Hemolymph osmolality (mOsm kg –1 ) Hemolymph osmolality (mOsm kg –1 )<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

(b) 1400<br />

1200<br />

1000<br />

H. crispae<br />

Osmotic performance<br />

Isoosmotic line<br />

0<br />

0 200 400 600 800 1000 1200 1400<br />

800<br />

600<br />

400<br />

200<br />

R. coronifer<br />

Osmotic performance<br />

Isoosmotic line<br />

0<br />

0 200 400 600 800 1000 1200 1400<br />

External Osmolality (mOsm kg –1 )<br />

Figure 6 Osmoregulation in eutardigrades. (a) Measured hemolymph<br />

osmolality of active stage Halobiotus crispae from<br />

Vellerup Vig, Denmark as a function of external osmolality.<br />

Figure modified from <strong>Halberg</strong> et al. (2009b). (b) Hemolymph<br />

osmolality of Richtersius coronifer from Öland, Sweden as a<br />

function of external osmolality. External solutions with osmotic<br />

concentrations of 100, 200, 300 and 500 mOsm kg )1 were<br />

prepared from artificial seawater salt (Tropic Marin, Dr Biener<br />

GmbH, Germany). Each point on the graph represents<br />

mean SD of hemolymph osmolality measurements (nanolitre<br />

osmometry; Clifton Technical Physics, Hartford, NY, USA)<br />

made in tardigrades kept in demineralized water and the different<br />

salt water solution for 30 min. Five animals were used for<br />

hemolymph determination at each of the experimental solutions.<br />

Arrow indicates the upper lethal line for R. coronifer.<br />

concentrations, R. coronifer will become inactive and<br />

eventually die, exhibiting an upper lethal limit of<br />

around 500 mOsm kg )1 (Fig. 6b). In solutions ranging<br />

from demineralized water to salt water with an osmolality<br />

of 500 mOsm kg )1 , the tardigrade, however,<br />

maintains a consistent osmotic gradient of around<br />

170 mOsm kg )1 above that of the external environment<br />

(Fig. 6b). In comparison, H. crispae exposed to the<br />

same salinity range maintains an osmotic gradient of<br />

around 300 mOsm kg )1 above that of the surroundings<br />

(Fig. 6a). These data on two different species of<br />

tardigrades indicate that eutardigrades have a relatively<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x 417


Adaptation to extreme environments in tardigrades Æ<br />

N. Møbjerg et al. Acta Physiol 2011, 202, 409–420<br />

high water turnover, and that they excrete a hypoosmotic<br />

fluid – the likely organs involved in this<br />

excretion being Malpighian tubules and the gut system<br />

(Fig. 4) (Møbjerg & Dahl 1996, <strong>Halberg</strong> et al. 2009b).<br />

At present the composition of tardigrade extracellular<br />

fluids (as well as intracellular fluids) are unknown.<br />

Obviously, information on the composition of the<br />

hemolymph is much needed for our understanding of<br />

tardigrade physiology and future discussions on how<br />

tardigrades regulate their body fluids and especially<br />

which osmolytes they are regulating. It is likely that the<br />

osmotic pressure obtained through hyperregulation is<br />

partly built by organic solutes, which in turn may act as<br />

bioprotectants or may be important for e.g. the above<br />

mentioned ability to supercool.<br />

In summary, recent research has shown that tardigrades<br />

in their active state may tolerate large fluctuations<br />

in abiotic factors and cope with environmental<br />

alterations by staying active and maintaining high<br />

metabolic rates. It would seem that these extraordinary<br />

organisms have two ways of handling environmental<br />

extremes; one way is entering a dormant state, i.e.<br />

cryptobiosis or diapause/encystment and the other,<br />

likely requiring a high metabolic rate, is relaying on<br />

e.g. osmoregulation and DNA repair, while staying<br />

active. The mechanisms enabling tardigrades to withstand<br />

environmental extremes in their active state may<br />

provide clues to the much debated mechanism underlying<br />

the phenomenon of cryptobiosis. Intriguingly,<br />

adaptations to some of the harshest environments on<br />

Earth have given tardigrades the ability to survive<br />

conditions that by far exceed the extremes presented by<br />

the environment (e.g. extreme levels of ionizing radiation<br />

and extremely low temperatures), presenting an<br />

unresolved puzzle for contemporary biology.<br />

Conflict of interest<br />

There are no conflicts of interest.<br />

We would like to thank Peter Westh for help in generating the<br />

DSC data presented in Figure 5. Special thanks are due to the<br />

organizers of the August Krogh Symposium 2010: Erik Hviid<br />

Larsen, Ylva Hellsten and Jørgen Wojtaszewski. Funding came<br />

from the 2008 Faculty of Science, University of Copenhagen<br />

Freja-Programme and from the Carlsberg Foundation.<br />

References<br />

Aguinaldo, A.M.A., Turbeville, J.M., Linford, L.S., Rivera,<br />

M.C., Garey, J.R., Raff, R.A. & Lake, J.A. 1997. Evidence<br />

for a clade of nematodes, arthropods and other moulting<br />

animals. Nature 387, 489–493.<br />

Arakawa, K., Ito, T., Kunieda, T., Horikawa, D., Soga, T. &<br />

Tomita, M. 2009. Comparative metabolome profiling of<br />

active and anhydrobiotic states of tardigrade Ramazzottius<br />

varieornatus. 11th International Symposium on Tardigrada.<br />

Tübingen, Germany. 3–6 August 2009, Conference Guide, p.<br />

37 (Abstract).<br />

Bavan, S., Straub, V.A., Blaxter, M.L. & Ennion, S.J. 2009. A<br />

P2X receptor from the tardigrade species Hypsibius dujardini<br />

with fast kinetics and sensitivity to zinc and copper.<br />

BMC Evol Biol 9, 17.<br />

Bertolani, R. 1970a. Mitosi somatiche e costanza cellulare<br />

numerica nei Tardigradi. Atti Accad Naz Lincei Rend Ser 8a,<br />

739–742.<br />

Bertolani, R. 1970b. Variabilità numerica cellulare in alcuni<br />

tessuti di Tardigradi. Atti Accad Naz Lincei Rend Ser 8a,<br />

442–445.<br />

Bertolani, R. 1982. Cytology and reproductive mechanisms in<br />

tardigrades. In: D.R. Nelson (ed.) Proceedings of the Third<br />

International Symposium on the Tardigrada, pp. 93–114.<br />

East Tennessee State University Press, Johnson City, TN.<br />

Bertolani, R., Guidetti, R., Jönsson, K.I., Altiero, T., Boschini,<br />

D. & Rebecchi, L. 2004. Experiences with dormancy in<br />

tardigrades. J Limnol 63, 16–25.<br />

Burnell, A.M., Houthoofd, K., O’Hanlon, K. & Vanfleteren, J.R.<br />

2005. Alternate metabolism during the dauer stage of the<br />

nematodeCaenorhabditiselegans.ExpGerontol40,850–856.<br />

Cassada, R.C. & Russell, R.L. 1975. The dauerlarva, a postembryonic<br />

developmental variant of the nematode Caenorhabditis<br />

elegans. Dev Biol 46, 326–342.<br />

Clegg, J.S. 1965. The origin of trehalose and its significance<br />

during formation of encysted dormant embryos of Artemia<br />

salina. Comp Biochem Physiol 14, 135–143.<br />

Clegg, J.S. 2001. Cryptobiosis – a peculiar state of biological<br />

organization. Comp Biochem Physiol B 128, 613–624.<br />

Crowe, J.H. 1972. Evaporative water loss by tardigrades under<br />

controlled relative humidity. Biol Bull 142, 407–416.<br />

Crowe, J.H. & Higgins, R.P. 1967. The revival of Macrobiotis<br />

aereolatus Murray (Tardigrada) from the cryptobiotic state.<br />

Trans Am Microsc Soc 86, 286–294.<br />

Crowe, J.H., Carpenter, J.F., Crowe, L.M. & Anchordoguy,<br />

T.J. 1990. Are Freezing and dehydration similar stress vectors<br />

– a comparison of modes of interaction of stabilizing<br />

solutes with biomolecules. Cryobiology 27, 219–231.<br />

Crowe, J.H., Hoekstra, F.A. & Crowe, L.M. 1992. Anhydrobiosis.<br />

Annu Rev Physiol 54, 579–599.<br />

Degma, P., Bertolani, R. & Guidetti, R. 2010. Actual checklist<br />

of Tardigrada species (Ver. 13: 01-06-2010). http://<br />

www.tardigrada.modena.unimo.it.<br />

Dewel, R.A. & Dewel, W.C. 1996. The brain of Echiniscus viridissimus<br />

Peterfi, 1956 (Heterotardigrada): a key to understanding<br />

the phylogenetic position of tardigrades and the<br />

evolution of the arthropod head. Zool J Linn Soc 116, 35–49.<br />

Dolezel, J., Bartoš, J., Voglmayr, H. & Greilhuber, J. 2003.<br />

Nuclear DNA content and genome size of trout and human.<br />

Cytometry 51A, 127–128.<br />

Dunn, C.W., Hejnol, A., Matus, D.Q., Pang, K., Browne, W.E.,<br />

Smith, S.A., Seaver, E., Rouse, G.W., Obst, M., Edgecombe,<br />

G.D. et al. 2008. Broad phylogenomic sampling improves<br />

resolution of the animal tree of life. Nature 452, 745–750.<br />

Edgecombe, G.D. 2010. Arthropod phylogeny: an overview<br />

from the perspective of morphology, molecular data and the<br />

fossil record. Arthropod Struct Dev 39, 74–87.<br />

418<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x


Acta Physiol 2011, 202, 409–420<br />

N. Møbjerg et al. Æ<br />

Adaptation to extreme environments in tardigrades<br />

Eibye-Jacobsen, J. 1997. Development, ultrastructure and<br />

function of the pharynx of Halobiotus crispae Kristensen,<br />

1982 (Eutardigrada). Acta Zool (Stock) 78, 329–347.<br />

Förster, F., Liang, C., Shkumatov, A., Beisser, D., Engelmann,<br />

J.C., Schnölzer, M., Frohme, M., Müller, T., Schill, R.O. &<br />

Dandekar, T. 2009. Tardigrade workbench: comparing<br />

stress-related proteins, sequence-similar and functional protein<br />

clusters as well as RNA elements in tardigrades. BMC<br />

Genomics 10, 469.<br />

Gabriel, W.N. & Goldstein, B. 2007. Segmental expression of<br />

Pax3/7 and Enrailed homologs in tardigrade development.<br />

Dev Genes Evol 217, 421–433.<br />

Goldstein, B. & Blaxter, M. 2002. Tardigrades. Curr Biol 12,<br />

R475.<br />

Goyal, K., Walton, L.J. & Tunnacliffe, A. 2005. LEA proteins<br />

prevent protein aggregation due to water stress. Biochem J<br />

388, 151–157.<br />

Gregory, T.R. 2010. Animal Genome Size Database. http://<br />

www.genomesize.com.<br />

Greven, H. 2007. Comments on the eyes of tardigrades.<br />

Arthropod Struct Dev 36, 401–407.<br />

Guidetti, R. & Bertolani, R. 2005. Tardigrade taxonomy: an<br />

updated check list of the taxa and a list of characters for their<br />

identification. Zootaxa 845, 1–46.<br />

Guidetti, R., Boschini, D., Altiero, T., Bertolani, R. & Rebecchi,<br />

L. 2008. Diapause in tardigrades: a study of factors involved<br />

in encystment. J Exp Biol 211, 2296–2302.<br />

<strong>Halberg</strong>, K.A., Persson, D., Møbjerg, N., Wanninger, A. &<br />

Kristensen, R.M. 2009a. Myoanatomy of the marine tardigrade<br />

Halobiotus crispae (Eutardigrada: Hypsibiidae).<br />

J Morph 270, 996–1013.<br />

<strong>Halberg</strong>, K.A., Persson, D., Ramløv, H., Westh, P., Kristensen,<br />

R.M. & Møbjerg, N. 2009b. Cyclomorphosis in Tardigrada:<br />

adaptation to environmental constraints. J Exp Biol 212,<br />

2803–2811.<br />

Hengherr, S., Heyer, A.G., Köhler, H.R. & Schill, R.O. 2008.<br />

Trehalose and anhydrobiosis in tardigrades – evidence for<br />

divergence in response to dehydration. FEBS 275, 281–288.<br />

Hengherr, S., Worland, M.R., Reuner, A., Brümmer, F. & Schill,<br />

R.O. 2009. Freeze tolerance, supercooling pionts and ice formation:<br />

comparative studies on the subzero temperature survival<br />

of limno-terrestrial tardigrades. J Exp Biol 212, 802–807.<br />

Hengherr, S., Reuner, A., Brümmer, F. & Schill, R.O. 2010. Ice<br />

crystallization and freeze tolerance in embryonic stages of<br />

the tardigrade Milnesium tardigradum. Comp Biochem<br />

Physiol A Mol Integr Physiol 156, 151–155.<br />

Horikawa, D.D., Sakashita, T., Katagiri, C., Watanabe, M.,<br />

Kikawada, T., Nakahara, Y., Hamada, N., Wada, S.,<br />

Funayama, T., Higashi, S., Kobayashi, Y., Okuda, T. &<br />

Kuwabara, M. 2006. Radiation tolerance in the tardigrade<br />

Milnesium tardigradum. Int J Radiat Biol 82, 843–848.<br />

Horikawa, D.D., Kuenida, T., Abe, W., Watanabe, M., Nakahara,<br />

Y., Yukuhiro, F., Sakashita, T., Hamada, N., Wada, S.,<br />

Funayama, T., Katagiri, C., Kobayashi, Y., Higashi, S. &<br />

Okuda, T. 2008. Establishment of a rearing system of the extremotolerant<br />

tardigrade Ramazzottius varieornatus: a new<br />

model animal for astrobiology. Astrobiology 8, 549–556.<br />

Jönsson, K.I. & Persson, O. 2010. Trehalose in three species of<br />

desiccation tolerant tardigrades. Open Zool J 3, 1–5.<br />

Jönsson, K.I. & Schill, R. 2007. Induction of Hsp70 by desiccation,<br />

ionising radiation and heat-shock in the eutardigrade<br />

Richtersius coronifer. Comp Biochem Physiol B 146,<br />

456–460.<br />

Jönsson, K.I., Borsari, S. & Rebecchi, L. 2001. Anhydrobiotic<br />

survival in populations of the tardigrades Richtersius coronifer<br />

and Ramazzottius oberhaeuseri from Italy and Sweden.<br />

Zool Anz 240, 419–423.<br />

Jönsson, K.I., Harms-Ringdahl, M. & Torudd, J. 2005. Radiation<br />

tolerance in the eutardigrade Richtersius coronifer. Int<br />

J Radiat Biol 81, 649–656.<br />

Jönsson, I.K., Rabbow, E., Schill, R.O., Harms-Ringdahl, M.<br />

& Rettberg, P. 2008. Tardigrades survive exposure to space<br />

in low Earth orbit. Curr Biol 18, R729–R731.<br />

Jørgensen, A., Faurby, S., Hansen, J.G., Møbjerg, N. & Kristensen,<br />

R.M. 2010. Molecular phylogeny of Arthrotardigrada<br />

(Tardigrada). Mol Phylogenet Evol 54, 1006–1015.<br />

Jørgensen, A. & Kristensen, R.M. 2004. Molecular phylogeny<br />

of Tardigrada – the monophyly of Heterotardigrada. Mol<br />

Phylogenet Evol 32, 666–670.<br />

Jørgensen, A., Møbjerg, N. & Kristensen, R.M. 1999. Ultrastructural<br />

studies on spermiogenesis and postcopulatory<br />

modifications of spermatozoa of Actinarctus doryphorus<br />

Schulz, 1935 (Arthrotardigrada: Halechiniscidae). Zool Anz<br />

238, 235–257.<br />

Jørgensen, A., Møbjerg, N. & Kristensen, R.M. 2007. A<br />

molecular study of the tardigrade Echiniscus testudo (Echiniscidae)<br />

reveals low DNA sequence diversity over a large<br />

geographical area. J Limnol 66(Suppl. 1), 77–83.<br />

Jørgensen, A., Møbjerg, N. & Kristensen, R.M. 2011. Phylogeny<br />

and evolution of the Echiniscidae (Echiniscoidea,<br />

Tardigrada) – an investigation of the congruence between<br />

molecules and morphology. J Zool Syst Evol Res: doi:<br />

10.1111/j.1439-0469.2010.00592.x<br />

Katayama, T., Arakawa, K., Hasebe, Y., Kido, N., Kunieda,<br />

T., Toyoda, A., Shin-I, T., Horikawa, D.D., Kuwahara, H.,<br />

Ohishi, K., Motoyama, A., Aizu, T., Kanehisa, M., Kohara,<br />

Y. & Fujiyama, A. 2009. Draft genome sequence assembly<br />

and preliminary annotations of Ramazzottius varieornatus<br />

genome. 11th International Symposium on Tardigrada.<br />

Tübingen, Germany. 3–6 August 2009, Conference Guide, p.<br />

36 (Abstract).<br />

Keilin, D. 1959. The problem of anabiosis or latent life: history<br />

and current concept. Proc Roy Soc Lond (Ser B) 150, 149–<br />

191.<br />

Kinchin, I.M. 1994. The Biology of Tardigrades, p. 186.<br />

Portland Press, London.<br />

Kristensen, R.M. 1982. The first record of cyclomorphosis in<br />

Tardigrada based on a new genus and species from Arctic<br />

meiobenthos. J Zool Syst Evol Res 20, 249–270.<br />

Kristensen, R.M. & Hallas, T.E. 1980. The tidal genus Echiniscoides<br />

and its variability with erection of Echiniscoididae<br />

fam.n. (Tardigrada). Zool Scr 9, 113–127.<br />

Maas, A. & Waloszek, D. 2001. Cambrian derivatives of the<br />

early arthropod stem lineage, pentastomids, tardigrades and<br />

lobopodians – an ‘Orsten’ perspective. Zool Anz 240, 451–<br />

459.<br />

Mali, B., Grohme, M.A., Förster, F., Dandekar, T., Schnölzer,<br />

M., Reuter, D., Wełnicz, W., Schill, R.O. & Frohme, M.<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x 419


Adaptation to extreme environments in tardigrades Æ<br />

N. Møbjerg et al. Acta Physiol 2011, 202, 409–420<br />

2010. Transcriptome survey of the anhydrobiotic tardigrade<br />

Milnesium tardigradum in comparison with Hypsibius<br />

dujardini and Richtersius coronifer. BMC Genomics 11,<br />

168.<br />

May, R.M., Maria, M. & Guimard, J. 1964. Actions différentielles<br />

des rayons x et ultraviolets sur le tardigrade Macrobiotus<br />

areolatus, àl¢état actif et desséché. Bull Biol Fr Belg<br />

98, 349–367.<br />

Møbjerg, N. & Dahl, C. 1996. Studies on the morphology and<br />

ultrastructure of the Malpighian tubules of Halobiotus<br />

crispae Kristensen, 1982 (Eutardigrada). Zool J Linn Soc<br />

116, 85–99.<br />

Møbjerg, N., Jørgensen, A., Eibye-Jacobsen, J., <strong>Halberg</strong>, K.A.,<br />

Persson, D. & Kristensen, R.M. 2007. New records on cyclomorphosis<br />

in the marine eutardigrade Halobiotus crispae.<br />

J Limnol 66(Suppl. 1), 132–140.<br />

Nelson, D.R. 2001. Tardigrada. In: J. Thorp & A. Covich (eds)<br />

Ecology and Classification of North American Freshwater<br />

Invertebrates, 2nd edn, pp. 527–550. Academic Press, San<br />

Diego.<br />

Nelson, D.R. 2002. Current status of the Tardigrada: evolution<br />

and ecology. Integr Comp Biol 42, 652–659.<br />

Neumann, S., Reuner, A., Brümmer, F. & Schill, R.O. 2009.<br />

DNA damage in storage cells of anhydrobiotic tardigrades.<br />

Comp Biochem Physiol A Mol Integr Physiol 153, 425–<br />

429.<br />

Nielsen, C. 2001. Animal Evolution: Interrelationships of<br />

the Living Phyla, 2nd edn. Oxford University Press,<br />

Oxford.<br />

Persson, D., <strong>Halberg</strong>, K.A., Jørgensen, A., Ricci, C., Møbjerg,<br />

N. & Kristensen, R.M. 2011. Extreme stress tolerance in<br />

tardigrades: surviving space conditions in low earth orbit.<br />

J Zoolog Syst Evol Res: doi: 10.1111/j.1439-0469.2010.<br />

00605.x.<br />

Ramazzotti, G. 1962. Il Phylum Tardigrada. Mem Insti Ital<br />

Idro Dott Marco De Marchi 14, 1–732.<br />

Ramazzotti, G. 1972. Tardigradi delle isole Kerguelen e descrizione<br />

della nouva specie Hypsibius (I) renaudi. Mem Ist<br />

Ital Idrobiol 29, 141–144.<br />

Ramazzotti, G. & Maucci, W. 1983. Il Phylum Tardigrada.<br />

Terza edizione riveduta e corretta. Mem Insti Ital Idro Dott<br />

Marco De Marchi 41, 1–1012.<br />

Ramløv, H. & Westh, P. 1992. Survival of the cryptobiotic<br />

tardigrade Adorybiotus coronifer during cooling to )196 °C:<br />

effect of cooling rate, trehalose level and short term preacclimation.<br />

Cryobiology 29, 125–130.<br />

Ramløv, H. & Westh, P. 2001. Cryptobiosis in the eutardigrade<br />

Adorybiotus (Richtersius) coronifer: tolerance to<br />

alcohols, temperature and de novo protein synthesis. Zool<br />

Anz 240, 517–523.<br />

Rebecchi, L. & Bertolani, R. 1994. Maturative patterne of<br />

ovary and testis in eutardigrades of fresh-water and terrestrial<br />

habitats. Invert Rep Dev 26, 107–117.<br />

Rebecchi, L., Altiero, T. & Guidetti, R. 2007. Anhydrobiosis:<br />

the extreme limit of desiccation tolerance. Invertebr Survival<br />

J 4, 65–81.<br />

Rebecchi, L., Altiero, T., Guidetti, R., Cesari, M., Bertolani,<br />

R., Negroni, M. & Rizzo, A.M. 2009a. Tardigrade resistance<br />

to space effects: first results of experiments on the<br />

LIFE-TARSE mission on FOTON-M3 (September 2007).<br />

Astrobiology 6, 581–591.<br />

Rebecchi, L., Cesari, M., Altiero, T., Frigieri, A. & Guidetti, R.<br />

2009b. Survival and DNA degradation in anhydrobiotic<br />

tardigrades. J Exp Biol 212, 4033–4039.<br />

Renaud-Mornant, J. 1982. Species diversity in marine Tardigrada.<br />

In: D.R. Nelson (ed.) Proceedings of the 3rd International<br />

Symposium on Tardigrada, pp. 149–178. East<br />

Tennessee State University Press, Johnson City.<br />

Reuner, A., Hengherr, S., Brahim, M., Förster, F., Arndt, D.,<br />

Reinhardt, R., Dandekar, T., Frohme, M., Brümmer, F. &<br />

Schill, R.O. 2010. Stress response in tardigrades: differential<br />

gene expression of molecular chaperones. Cell Stress Chaperones<br />

15, 423–430.<br />

Rizzo, A.M., Negroni, M., Altiero, T., Montorfano, G.,<br />

Corsetto, P., Berselli, P., Berra, B., Guidetti, R. & Rebecchi,<br />

L. 2010. Antioxidant defences in hydrated and desiccated<br />

states of the tardigrade Paramacrobiotus richtersi. Comp<br />

Biochem Physiol B 156, 115–121.<br />

Sands, C.J., McInnes, S.J., Marley, N.J., Goodall-Copestake,<br />

W.P., Convey, P. & Linse, K. 2008. Phylum Tardigrada: an<br />

‘‘individual’’ approach. Cladistics 24, 1–11.<br />

Schill, R.O. 2010. Anhydrobiotic abilities of tardigrades. In:<br />

S. Hohmann (Series ed.) Topics in Current Genetics Vol. 21,<br />

E. Lubzens, J. Cerda & M.S. Clark (eds) Dormancy and<br />

resistance in harsh environments, pp. 133–146, Springer-<br />

Verlag, Heidelberg.<br />

Schill, R.O., Steinbrück, G.H.B. & Köhler, H.R. 2004. Stress<br />

gene (hsp 70) and quantitative expression in Milnesium<br />

tardigradum (Tardigrada) during active and cryptobiotic<br />

stages. J Exp Biol 207, 1607–1613.<br />

Schokraie, E., Hotz-Wagenblatt, A., Warnken, U., Mali, B.,<br />

Frohme, M., Förster, F., Dandekar, T., Hengherr, S., Schill,<br />

R.O. & Schnölzer, M. 2010. Proteomic analysis of tardigrades:<br />

towards a better understanding of molecular mechanisms<br />

by anhydrobiotic organisms. PLoS One 5, e9502.<br />

Suzuki, A.C. 2003. Life history of Milnesium tardigradum<br />

Doyère (Tardigrada) under a rearing environment. Zool Sci<br />

20, 49–57.<br />

Warner, A.H., Miroshnychenko, O., Kozarova, A., Vacratsis,<br />

P.O., MacRae, T.H., Kim, J. & Clegg, J.S. 2010. Evidence<br />

for multiple group 1 late embryogenesis abundant (LEA)<br />

proteins in encysted embryos of Artemia and their organelles.<br />

J Biochem 148, 581–592.<br />

Westh, P. & Kristensen, R.M. 1992. Ice formation in the<br />

freeze-tolerant eutardigrades Adorybiotus coronifer and<br />

Amphibolus nebulosus studied by differential scanning calorimetry.<br />

Polar Biol 12, 693–699.<br />

Westh, P. & Ramløv, H. 1991. Trehalose accumulation in the<br />

tardigrade Adorybiotus coronifer during anhydrobiosis.<br />

J Exp Zool 258, 303–311.<br />

Westh, P., Kristiansen, J. & Hvidt, A. 1991. Ice-nucleating<br />

activity in the freezetolerant tardigrade Adorybiotus coronifer.<br />

Comp Biochem Physiol A 99, 401–404.<br />

Wright, J.C. 2001. Cryptobiosis 300 years on from van Leuwenhoek:<br />

what have we learned about tardigrades Zool<br />

Anz 240, 563–582.<br />

Wright, J.C., Westh, P. & Ramløv, H. 1992. Cryptobiosis in<br />

Tardigrada. Biol Rev 67, 1–29.<br />

420<br />

Ó 2011 The Authors<br />

Acta Physiologica Ó 2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02252.x


Paper IV


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

RESEARCH ARTICLE<br />

Research article<br />

Functional characterization of the vertebrate<br />

primary ureter: Structure and ion transport<br />

mechanisms of the pronephric duct in axolotl<br />

larvae (Amphibia)<br />

Birgitte M Haugan, <strong>Kenneth</strong> A <strong>Halberg</strong>, Åse Jespersen, Lea R Prehn and Nadja Møbjerg*<br />

Open Access<br />

Abstract<br />

Background: Three kidney systems appear during vertebrate development: the pronephroi, mesonephroi and<br />

metanephroi. The pronephric duct is the first or primary ureter of these kidney systems. Its role as a key player in the<br />

induction of nephrogenic mesenchyme is well established. Here we investigate whether the duct is involved in urine<br />

modification using larvae of the freshwater amphibian Ambystoma mexicanum (axolotl) as model.<br />

Results: We investigated structural as well as physiological properties of the pronephric duct. The key elements of our<br />

methodology were: using histology, light and transmission electron microscopy as well as confocal laser scanning<br />

microscopy on fixed tissue and applying the microperfusion technique on isolated pronephric ducts in combination<br />

with single cell microelectrode impalements. Our data show that the fully differentiated pronephric duct is composed<br />

of a single layered epithelium consisting of one cell type comparable to the principal cell of the renal collecting duct<br />

system. The cells are characterized by a prominent basolateral labyrinth and a relatively smooth apical surface with one<br />

central cilium. Cellular impalements demonstrate the presence of apical Na + and K + conductances, as well as a large K +<br />

conductance in the basolateral cell membrane. Immunolabeling experiments indicate heavy expression of Na + /K + -<br />

ATPase in the basolateral labyrinth.<br />

Conclusions: We propose that the pronephric duct is important for the subsequent modification of urine produced by<br />

the pronephros. Our results indicate that it reabsorbs sodium and secretes potassium via channels present in the apical<br />

cell membrane with the driving force for ion movement provided by the Na + /K + pump. This is to our knowledge the<br />

first characterization of the pronephric duct, the precursor of the collecting duct system, which provides a model of<br />

cell structure and basic mechanisms for ion transport. Such information may be important in understanding the<br />

evolution of vertebrate kidney systems and human diseases associated with congenital malformations.<br />

Background<br />

During the development from embryo to adult life vertebrates<br />

use a succession of kidney forms to maintain extracellular<br />

fluid homeostasis and simultaneously rid the<br />

body of nitrogenous wastes [1,2]. Three spatially and<br />

temporally different kidney generations form from the<br />

intermediate mesoderm in an anterior to posterior direction<br />

i.e. the pronephroi, mesonephroi and metanephroi<br />

[3]. The functional unit in these paired kidneys is the<br />

* Correspondence: nmobjerg@bio.ku.dk<br />

1 Department of Biology, University of Copenhagen, Universitetsparken, DK-<br />

2100 Copenhagen, Denmark<br />

Full list of author information is available at the end of the article<br />

nephron, and it is composed of a filtration unit and a<br />

renal tubule. Urine is produced by the filtration of blood<br />

in the filtration unit, followed by the selective reabsorption<br />

and secretion of ions, organic molecules and water<br />

across highly specialized epithelia of the renal tubule [4].<br />

Nephrons open into a ureter, and in the meso- and metanephros<br />

they do so via a collecting duct system [5,6]. This<br />

system is the site for the important final adjustment of<br />

the urine. Vertebrate kidneys may produce urine, which is<br />

either hypoosmotic (diluted), isoosmotic or hyperosmotic<br />

(concentrated) relative to the body fluids [7-10].<br />

This ability is a function of i) the evolutionary state of the<br />

© 2010 Haugan et al; licensee BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons<br />

Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in<br />

any medium, provided the original work is properly cited.


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 2 of 9<br />

nephrons and ii) the regulation of filtration and of the<br />

transport of inorganic ions, organic molecules and water<br />

across the renal epithelia.<br />

The first kidneys to form - the embryonic pronephric<br />

kidneys - are the functional kidneys of fish and amphibian<br />

larvae [11-16]. These are very simple kidneys composed<br />

of a single nephron. A characteristic of vertebrate<br />

kidney organogenesis is the development of a pronephric<br />

duct in association with each pronephros [3]. These ducts<br />

are the first or primary ureters of vertebrate kidney systems.<br />

They form the collecting duct system of the mesoand<br />

metanephroi, and they, and their derivates, are the<br />

key players in the induction of the nephrogenic mesenchyme,<br />

which forms these latter kidney generations. Few<br />

functional studies exist on the pronephros and functional<br />

studies of the duct are virtually lacking. Molecular studies<br />

have been directed at mapping genes expressed in different<br />

segments of the pronephric nephron, and several<br />

recent reviews have highlighted the potential use of this<br />

embryonic kidney in drug and human kidney disease<br />

assessment [17-23]. To date, there has been little focus on<br />

the role of the pronephric duct in urine modification and<br />

it remains to be shown whether transepithelial transport<br />

processes are present in this structure. Ultrastructural<br />

investigations have shown duct cells with the characteristics<br />

of an epithelium involved in active transport e.g.<br />

many mitochondria and surface expansions of the basolateral<br />

cell membranes [14,24]. In addition, gene expression<br />

assays have indicated high expression of transporters<br />

known to be involved in ion transport, such as the Na + /<br />

K + -ATPase and the ROMK channel [17,18,25-27]. Collectively,<br />

these data suggest that the pronephric duct may<br />

play an important role in regulation of extracellular fluid<br />

homeostasis. Therefore we ask the question: Is the duct<br />

involved in urine modification<br />

In amphibians the pronephros is a large organ, which is<br />

functional for a considerable time, before it degenerates.<br />

We investigate structural and functional characteristics of<br />

the pronephric duct in the freshwater amphibian Ambystoma<br />

mexicanum (axolotl). Members of the Ambystoma<br />

genus have been used as models for the study of pronephric<br />

structure, function, development and evolution<br />

for more than a century [24,28,29] and the formation and<br />

caudal migration of the pronephric duct in the axolotl has<br />

been thoroughly investigated [30-32]. Numerous functional<br />

studies exist on the mesonephric collecting duct<br />

system of both urodele and anuran amphibians, which<br />

provide detailed information on the transport characteristics<br />

of these segments [Reviewed in [13]]. Our histological<br />

examinations and dissections of axolotl larvae<br />

indicate that the pronephros is functional from the time<br />

of hatching to larval stage 54. We investigate duct morphology<br />

and cellular transport mechanisms present in<br />

larvae with functional pronephroi, and show that the primary<br />

ureter is important for urine dilution in the axolotl.<br />

The single cell type found in the ureter shows the characteristics<br />

of the vertebrate collecting duct system principal<br />

cell and our data indicate that it reabsorbs Na + and<br />

secretes K + .<br />

Results and Discussion<br />

Identification of functional pronephroi and pronephric<br />

ducts<br />

We determined the interval in which the axolotl pronephros<br />

and pronephric duct is functional by investigating<br />

kidney structure in freshly dissected larvae and in<br />

larvae prepared for histology (Figure 1 and 2). The pronephroi<br />

of axolotl larvae are paired organs located on<br />

each side of the dorsal aorta in the most anterior part of<br />

the body cavity. They are visible from the outside on the<br />

dorsal side of the larva as two small bulges behind the<br />

gills (Figure 1A). Each of the two kidneys are composed of<br />

a filtrating unit - a glomus originating from the dorsal<br />

aorta, and a single convoluted renal tubule opening into<br />

the coelom via two ciliated nephrostomes (Figure 1B).<br />

The pronephros is fully functional when the axolotl larva<br />

Figure 1 Axolotl larva stage 45 and 52. A. Dorsal view of recently<br />

hatched axolotl larva (stage 45; forelimb present as limb buds). The<br />

pronephroi can be seen as two small bulges behind the gills. B. Schematic<br />

representation of the pronephric nephron in the stage 52 larva<br />

as revealed from light- and transmission electron microscopy on serial<br />

section of plastic embedded tissue. The pronephros consists of an external<br />

glomus (gl) and a single renal tubule, which opens into the<br />

coelom (co) via two ciliated nephrostomes (ne). In this late larval stage<br />

the tubule is divided into two ciliated tubules (ci), two proximal tubule<br />

branches (pt), a common proximal tubule, a ciliated intermediate segment<br />

(is) and a distal tubule (dt). The distal tubule continues as the pronephric<br />

duct (pd), which leaves the confines of the pronephros and<br />

empties into the cloaca.


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 3 of 9<br />

A<br />

pt<br />

dt<br />

nc<br />

ne<br />

gl<br />

is<br />

25 m<br />

co<br />

B<br />

50 m<br />

C<br />

pt<br />

mr<br />

wd<br />

mv<br />

pd<br />

co<br />

bl<br />

nu<br />

mu<br />

15m<br />

5 m<br />

Figure 2 Histology of pronephros and pronephric duct. A. Cross section of a stage 54 larva (forelimb completely developed) revealing the filtration<br />

unit and the convoluted pronephric tubule. Araldite section, 1.5 μm, stained with toluidine blue. Blood is filtered in the external glomus (gl) and the<br />

filtrate enters the coelom (co) before it is taken up into the renal tubule via ciliated nephrostomes (ne). In this late larval stage the tubule is characterized<br />

by possessing a ciliated intermediate segment (is). nc, notochord; dt, distal tubule; pt, proximal tubule. B. Longitudinal section of pronephric duct<br />

(stage 52 larva). Araldite section, 2 μm, stained with toluidine blue. The pronephric duct (pd) leaves the confines of the pronephros. co, coelom; mu,<br />

muscle; pt, proximal tubule. INSERT: The Wolffian duct (wd) at the level of the caudal part of the mesonephros in a stage 54 larva. The duct epithelium<br />

consists of two cell types: principal cells and intercalated, mitochondria-rich cells (mr). C. Transmission electron microscopy of pronephric duct shown<br />

in figure 2B. The duct is composed of a single cell type characterized by a relative smooth apical surface with few microvilli (mv) and a well developed<br />

basal labyrinth (bl) formed by the highly invaginated basal and to some extent lateral cell membranes. nu, nucleus.


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 4 of 9<br />

hatches from the egg at stage 44. This was confirmed by<br />

the appearance of blood cells in the capillaries of the glomus,<br />

which from embryonic stage 36 had a fully developed<br />

endothelium and a visceral layer with podocytes.<br />

From stage 44 the cilia of the nephrostomes points<br />

toward the lumen of the renal tubule, indicating a passage<br />

of fluid from the coelom. At this stage the kidney consists<br />

of an external glomus and a renal tubule with the following<br />

morphologically determined segments: two nephrostomes,<br />

each connected to a branch of proximal tubule, a<br />

common proximal tubule and a distal tubule. The pronephric<br />

duct runs caudally as an extension of the distal<br />

tubule opening into the cloaca. From the time of hatching<br />

the kidney was observed to increase in overall size due to<br />

further segmentation of the renal tubule i.e. the development<br />

of a ciliated intermediary segment; present from<br />

stage 52 (Figure 1B and 2A). The fully segmented pronephric<br />

tubule consists of the following morphological<br />

defined segments: two nephrostomial tubules, two proximal<br />

tubule branches, a common proximal tubule, a ciliated<br />

intermediary segment and a distal tubule (Figure<br />

2A). At stage 52 the mesonephros was clearly visible and<br />

functional as judged by glomerular maturation and presence<br />

of blood cells in the mesonephric glomerular capillaries.<br />

Hence, comparable to the situation found in<br />

anuran amphibians [14,33] the two kidney generations<br />

functionally overlap in axolotl larvae. At stage 54 the pronephros<br />

reaches its maximal size. Gonadal primordia<br />

were observed medioventral to the mesonephros (not<br />

shown). These were undifferentiated and sex determination<br />

was not possible. During the transition from stage 54<br />

to latter stages, characterized by fully developed hind<br />

limbs, overall pronephric tubule and glomus volume<br />

decreased, marking pronephric degeneration. Pronephric<br />

structure in stage 52-54 larvae of Ambystoma mexicanum<br />

resembled the description by Christensen (1964) of the<br />

functional pronephros in A. punctatum [24].<br />

Pronephric duct structure<br />

Structural examination of the pronephric duct in larvae<br />

with functional pronephroi revealed that the duct consists<br />

of a single cell type (Figure 2B and 2C). Thus, the<br />

heterocellularity with intercalated mitochondria-rich<br />

cells interposed between principal cells, characteristic of<br />

the collecting duct system of latter kidney generations, is<br />

not seen at this point (Figure 2B and 2C). This is comparable<br />

to the situation in the green toad, Bufo viridis - a<br />

terrestrial anuran amphibian [14]. The pronephric duct<br />

cells in A. mexicanum are approximately 20 μm high in<br />

early larval stages, but decrease in height to 10-15 μm in<br />

stage 52 and 54 larvae (Figure 2C). They have a relatively<br />

smooth apical surface. A single central cilium is present<br />

(not shown) in addition to small and sparse microvilli,<br />

which are more numerous at the point of the apical cell<br />

junctions. The nucleus is regularly shaped, centrally<br />

placed and contains a nucleolus and patches of heterochromatin.<br />

There is a conspicuous basal labyrinth, and<br />

lateral infoldings are seen as well (Figure 2C). The cytoplasm<br />

contains many mitochondria in addition to a Golgi<br />

complex and endoplasmatic reticulum. The morphology<br />

of the duct changes at the level of the mesonephros in late<br />

larval stages 52-54, revealing the presence of intercalated,<br />

mitochondria-rich cells (Figure 2B, insert).<br />

Ion transport mechanisms in the pronephric duct of larvae<br />

stage 46-54<br />

We examined if the pronephric duct participates in final<br />

urine modification with the aid of glass microelectrodes<br />

and ion substitution experiments in isolated and perfused<br />

ducts dissected from 22 larvae. Figure 3A is a frequency<br />

distribution of the membrane potential (V m ) of 64<br />

impaled cells. The data show a broad distribution with an<br />

Number of cells<br />

V m (mV)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

-100<br />

A<br />

0<br />

-110 -90 -70 -50 -30<br />

V m (mV)<br />

B<br />

high K<br />

2 min<br />

V m (mV)<br />

-40<br />

-60<br />

-80<br />

-100<br />

high K<br />

Figure 3 Voltage recordings from single cells in isolated and perfused<br />

ducts. A. Frequency distribution of the membrane potential<br />

(V m ) in 64 cells from pronephric ducts dissected from axolotl larvae in<br />

the stage 46-54. B. Original voltage trace from single cell. Effect of raising<br />

bath [K + ] from 3 to 20 mmol/l. V m depolarized indicating the presence<br />

of a basolateral K + conductance. C. Summary data illustrating the<br />

effect on V m of the bath K + concentration step (n = 29).<br />

0<br />

-20<br />

C<br />

-10


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 5 of 9<br />

average V m between -75 and -80 mV. Transport characteristics<br />

of the duct did not seem to differ between early<br />

and late larval stages. As shown in figure 3B and 3C, raising<br />

the K + concentration in the basal solution from 3 to<br />

20 mmol/l resulted in a large, reversible depolarization of<br />

V m , revealing the presence of a large basolateral K + conductance.<br />

This indicates the presence of K + channels in<br />

the basolateral cell membrane. We examined whether the<br />

duct has luminal electrogenic transporters or channels in<br />

experiments with luminal K + and Na + steps. Figure 4<br />

V m (mV)<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

A<br />

low Na<br />

high K<br />

low Na<br />

shows the result of these substitution experiments. V m<br />

hyperpolarized upon a decrease in luminal Na + concentration<br />

from 102 to 7 mmol/l and depolarized upon an<br />

increase in luminal K + concentration from 3 to 20 mmol/<br />

l. This indicates that the luminal (apical) cell membrane<br />

possesses Na + as well as K + channels. In order to identify<br />

an ion pump, which can provide the driving force for<br />

luminal uptake of Na + as well as K + secretion, we isolated<br />

pronephric ducts and performed immunolabeling with<br />

an antibody directed against the Na + -K + -ATPase α-subunit.<br />

As shown in figure 5, the Na + -K + -ATPase is highly<br />

expressed in the pronephric ducts from these larvae, and<br />

is entirely localized to the lateral and highly invaginated<br />

basolateral cell membranes.<br />

Electrophysiological studies performed on the mesonephric<br />

collecting duct system of amphibians, have indicated<br />

that principal cells in aquatic urodeles, have a large<br />

apical Na + conductance and no, or very small, K + conductance.<br />

However, in terrestrial anurans, K + secretion<br />

through apical K + channels seems a major task of the<br />

principal cells [34-39]. In the current study we provide<br />

evidence for a principal cell, which has the characteristics<br />

of the mammalian collecting duct principal cell, i.e. with<br />

luminal Na + as well as K + conductances.<br />

-100<br />

5 min<br />

A<br />

0<br />

B<br />

-20 low Na -20 high K<br />

0<br />

C<br />

V m (mV)<br />

-40<br />

-60<br />

V m (mV)<br />

-40<br />

-60<br />

B<br />

25ìm<br />

-80<br />

-80<br />

-100<br />

-100<br />

Figure 4 Electrophysiological response to luminal fluid exchange.<br />

A. Original voltage trace from single cell. Effect of changing<br />

luminal [Na + ] from 102 to 7 mmol/l and luminal [K + ] from 3 to 20<br />

mmol/l. V m hyperpolarized upon the decrease in luminal [Na + ] and depolarized<br />

upon an increase in luminal [K + ]. This indicates that the luminal<br />

(apical) cell membrane possesses Na + as well as K + conductances.<br />

B. Summary data illustrating the effect on V m of concentration steps in<br />

luminal Na + (n = 6) and K + (n = 5).<br />

DAPI<br />

75ìm + +<br />

Na -K -ATPase<br />

Figure 5 Na + -K + -ATPase expression. A. The Na + -K + -ATPase is highly<br />

expressed in the pronephric ducts and entirely localized to the lateral<br />

and highly invaginated basolateral cell membranes. B. Na + -K + -ATPase<br />

expression in duct counterstained with DAPI. Images are three-dimensional<br />

reconstructions of the original CLSM z-series, showing a median<br />

longitudinal section of the pronephric duct.


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 6 of 9<br />

Conclusions<br />

We show that the pronephric duct, which is the first or<br />

primary ureter in all vertebrates, participates in urine<br />

adjustment in the axolotl. The cells constituting the duct<br />

are on the ultrastructural as well as cell physiological level<br />

comparable to principal cells found in the first segments<br />

of the mammalian collecting duct system [5,6,14,40-42].<br />

Notably, the pronephric duct lacks intercalated, mitochondria-rich<br />

cells. We propose that the duct is important<br />

for urine dilution through NaCl reabsorbtion, and<br />

that it in addition participates in the regulation of K +<br />

homeostasis. Figure 6 provides a model of the ion transport<br />

mechanisms, which we suggest are present in the<br />

duct cell. In this model a Na + -K + -ATPase in the basolateral<br />

cell membrane pumps Na + out of the cell and thereby<br />

provides the driving force for apical uptake of Na +<br />

through channels. The epithelial sodium channel (ENaC)<br />

is a likely candidate mediating this apical Na + uptake [43-<br />

45]. Na + exits the cell through the pump. K + is secreted<br />

through apical channels, and is recycled for the pump<br />

across the basolateral cell membrane. It is highly probable<br />

that ROMK channels, known to mediate K + secretion in<br />

the mammalian collecting duct system [42], and shown to<br />

be expressed in amphibian pronephric ducts [18], mediate<br />

K + secretion across the apical cell membrane. The<br />

active transport of Na + would create a lumen-negative<br />

transepithelial potential, and Cl - would presumably follow<br />

passively through the paracellular pathway of this epithelium<br />

[45]. This is to our knowledge the first characterization<br />

of the pronephric duct-the precursor of the<br />

+<br />

Na<br />

+<br />

Na<br />

+<br />

+<br />

K K<br />

ATP<br />

Figure 6 Suggested model for ion transport mechanisms in the<br />

cells of the vertebrate primary ureter. A Na + -K + -ATPase in the basolateral<br />

cell membrane pumps Na + out of the cell and thereby provides<br />

the driving force for apical uptake of Na + through channels. K + is secreted<br />

across the apical cell membrane through channels and recycled for<br />

the pump across the basolateral cell membrane.<br />

collecting duct system- which provides detailed information<br />

on cell structure and the basic mechanisms for ion<br />

transport.<br />

Methods<br />

Animals<br />

Specimens of the Mexican axolotl Ambystoma mexicanum<br />

(Shaw and Nodder, 1798) came from the animal stable<br />

of the August Krogh Building, part of the Campus<br />

Animal Research Facility at University of Copenhagen.<br />

Staging were performed according to [46] for embryos<br />

and the larvae were designated according to the degree of<br />

limb development as defined by the Ambystoma Genetic<br />

Stock Center, University of Kentucky; http://www.ambystoma.org.<br />

Larvae used for experiments, were in the stage<br />

44 to 54. They were euthanized by decapitation, followed<br />

by brain destruction, and were subsequently either prepared<br />

for histology or pronephric ducts (o.d. 50-70 μm,<br />

dissected length 300-1000 μm) were free hand dissected<br />

at 6°C and prepared for microperfusion experiments or<br />

immunolabeling. The pronephric ducts were dissected<br />

from the region in front of the mesonephros. Dissections<br />

were performed in media containing (in mmol/l): 75<br />

NaCl, 20 NaHCO 3 , 3.0 KCl, 1.8 CaCl 2 , 1.0 MgSO 4 , 0.8<br />

Na 2 HPO 4 , 0.2 NaH 2 PO 4 , 5.5 glucose, 3.3 glycine, 0.4 PVP,<br />

5.0 HEPES, titrated to pH 7.8 with NaOH.<br />

Histology, Light and Transmission Electron Microscopy<br />

Dissections were performed for every developmental<br />

stage of post-hatched axolotls from stage 44 to 57. In<br />

addition, we examined pronephric development in<br />

embryos. Light microscopic imaging of live and dissected<br />

larvae was performed using Zeiss Stemi 2000-CS and<br />

Leica MZ 16 microscopes equipped with an Infinity X<br />

Digital Camera (DeltaPix, Denmark). For histology, a<br />

total of 31 larvae in the stage 44-54 and 19 embryos from<br />

stage 21 to 44 were used. Specimens were fixed for 12<br />

hours at room temperature in an aldehyde fixative containing:<br />

1.2% glutaraldehyde, 1% paraformaldehyde, 0.05<br />

mol/l sucrose and 0.05 mol/l sodium cacodylate buffer<br />

(pH 7.4) and subsequently rinsed and stored in 0.05 mol/l<br />

sodium cacodylate buffer with 0.05 mol/l sucrose. Following<br />

1 hour's post fixation in 2% OsO 4 with 0.1 mol/l<br />

sodium cacodylate, specimens were dehydrated through a<br />

graded series of ethanol and propylenoxide and embedded<br />

in Araldite. For light microscopy, 1.5 μm sections<br />

were cut with glass knives on a Leica ultramicrotome EM<br />

UC6 and stained with toluidine blue. Ultrathin sections<br />

for transmission electron microscopy were cut on the<br />

same microtome with a Diatome diamond knife and subsequently<br />

stained with uranyl acetate and lead citrate.<br />

Transmission electron microscopic images were acquired<br />

using JEOL 100SX and JEOL JEM 1011 transmission electron<br />

microscopes. Kodak negatives obtained from the


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 7 of 9<br />

JEOL 100SX were digitized using an Epson Perfection<br />

4990 Photo scanner. The JEOL JEM-1011 was equipped<br />

with a GATAN digital camera. Digital images were optimized<br />

for contrast and color using CorelDraw X4.<br />

Microperfusion and cellular impalements<br />

Pronephric ducts were transferred to a bath chamber<br />

mounted on an inverted microscope and perfused in vitro<br />

with a set of pipettes made to fit the diameter of the<br />

tubules [37-39]. The tubule perfusion system used (Luigs<br />

& Neumann, Germany) was designed for accurate adjustment<br />

and movement of concentric pipettes [47]. Holding<br />

and perfusion pipettes were hand made from glass tubing<br />

(Drummond Scientific Company, PA, USA; holding<br />

pipettes: o.d. 2.1 mm, i.d. 1.6 mm; perfusion pipettes: o.d.<br />

1.2 mm, i.d. 1.0 mm) in a microforge equipped with a<br />

microscope (SM II/1 Puller from Luigs & Neumann, Germany).<br />

The tubules were perfused with a single-barrelled<br />

perfusion pipette containing a small glass capillary (o.d.<br />

0.3 mm, i.d. 0.2 mm, Drummond Scientific Company, PA,<br />

USA) connected to a manual Hamilton valve (Hamilton<br />

Co., NV, USA) with a four-way distribution system. Fluid<br />

exchange during luminal perfusion experiments was<br />

made through this capillary, the tip of the capillary being<br />

placed close to the opening of the perfusion pipette,<br />

ensuring fast fluid exchange in the tubule. Luminal perfusion<br />

was performed either by hand through a syringe<br />

connected to one of the ports in the valve or it was gravity-driven<br />

through a port connected to a fluid filled reservoir.<br />

The pressure applied was adjusted by monitoring<br />

tubule diameter, ensuring that the tubule neither collapsed<br />

nor over expanded. The bath was perfused at 8 ml/<br />

min and fluid exchange was performed through beakers<br />

attached to the outside of the Faraday cage, which surrounds<br />

the microperfusion set-up.<br />

Perfusions of tubule bath and luminal fluid were carried<br />

out at room temperature with a perfusion solution containing<br />

(in mmol/l): 75 NaCl, 25 NaHCO 3 , 3.0 KCl, 1.8<br />

CaCl 2 , 1.0 MgSO 4 , 0.8 Na 2 HPO 4 , 0.2 NaH 2 PO 4 . The perfusion<br />

solution was equilibrated with 1.8% CO 2 in O 2 and<br />

had a measured pH of 7.8. Experimental solutions with<br />

different sodium and potassium concentrations were prepared<br />

from this control solution. In the high K + solution,<br />

the K + concentration was raised to 20 mmol/l by equimolar<br />

substitution with Na + . In the low Na + solution with a<br />

[Na + ] of 6.8 mmol/l, Na + was replaced by choline or N-<br />

methyl-D-glucamine (NMDG + ) titrated with HCl.<br />

Pronephric duct cells were impaled across the basal cell<br />

membrane with KCl (1-3 mol/l) filled glass microelectrodes<br />

(R electrode ≈ 100 MΩ) and the membrane potential<br />

(V m ) was recorded with respect to the grounded bath.<br />

The microelectrodes were pulled from borosilicate glass<br />

with filament (Clark Electromedical, UK) on a vertical<br />

electrode puller (Narishige, Japan). Impalements were<br />

achieved by placing the microelectrode tip against the<br />

basal surface of the cell and gently taping the micromanipulator<br />

(Leitz, Germany) on which the electrodes were<br />

mounted. Voltage recordings were made by a WPI Duo<br />

773 electrometer (World Precision Instruments, USA)<br />

and digitized by a PowerLab/4S data acquisition system<br />

(ADInstruments, Australia). The recording of V m was<br />

accepted if the impalement was achieved by a sudden<br />

change in the potential read by the electrode and if the<br />

impalement was stable.<br />

The results are based on 64 cell impalements made in<br />

22 pronephric ducts dissected from 22 axolotl larvae in<br />

the stage 46-54. Figures were made in Origin 7.5 (Microcal,<br />

USA) and CorelDRAW X4.<br />

Immunolabeling, Confocal Laser Scanning Microscopy and<br />

3D reconstruction<br />

For identification and localization of the Na + -K + -ATPase,<br />

three separate immunolabeling experiments were conducted<br />

with equal results covering larvae in stages 46-54.<br />

In each experiment, pronephric ducts were isolated from<br />

four to five specimens and subsequently fixed on ice for<br />

approximately 60 minutes in 3% paraformaldehyde buffered<br />

to pH 7.4 with 0.1 mol/l sodium cacodylate. After<br />

being rinsed in PBS (perfusion solution), the tissue was<br />

incubated overnight at 4°C in PBS containing 10% goat<br />

serum (Invitrogen, CA, USA), 1% triton-X and the Na + -<br />

K + ATPase monoclonal mouse antibody α5-IgG (10 μg/<br />

ml). The α5 antibody developed by D.M. Fambrough was<br />

obtained from the Developmental Studies Hybridoma<br />

Bank developed under the auspices of the NICHD and<br />

maintained by The University of Iowa, Department of<br />

Biology (Iowa City, IA 52242). Following an extensive<br />

wash in PBS, the pronephric ducts were incubated with<br />

Alexa-488-conjugated goat anti-mouse IgG (1:100, Invitrogen,<br />

CA, USA) overnight at 4°C. Following rinses in<br />

PBS, the tissue was mounted on glass coverslips in Vecta<br />

shield (Vector Laboratories Inc., CA, USA). In some<br />

preparations, the renal tubules were incubated in DAPI<br />

(1:250, Invitrogen, CA, USA) for approximately 5 min<br />

and washed in PBS, prior to mounting. Image acquisition<br />

was performed on a Leica DM RXE 6 TL inverted microscope<br />

equipped with a Leica TCS SP2 AOBS confocal<br />

laser scanning unit, using the 488 nm line of an argon/<br />

crypton laser. The image series were processed and<br />

edited in the 3-D reconstruction software IMARIS (Bitplane<br />

AG, Zürich, Switzerland). The confocal images are<br />

based on 150-170 optical sections of a Z-series performed<br />

at intervals of 0.3-0.5 μm. All control preparations<br />

were negative for immunostaining.<br />

Abbreviations<br />

bl: basal labyrinth; ci: ciliated tubule; co: coelom; dt: distal tubule; gl: glomus; is:<br />

ciliated intermediate segment; nc: notochord; ne: nephrostome; nu: nucleus;


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 8 of 9<br />

mr: intercalated mitochondria-rich cell; mu: muscle; mv: microvilli; pd: pronephric<br />

duct; pt: proximal tubule; V m : membrane potential; wd: Wolffian duct.<br />

Authors' contributions<br />

NM conceived and designed the study. BMH and NM fixed larvae and performed<br />

the dissections for the structural investigation. BMH sectioned specimens,<br />

and made LM and TEM investigations with help from ÅJ and NM. NM<br />

made the microperfusion experiments and microelectrode impalements. KAH,<br />

LRP and NM carried out immunostaining experiments and KAH performed<br />

CLSM and prepared the 3D images. All authors participated in discussions and<br />

interpretation of the data. NM wrote the paper with inputs from the other<br />

authors. All authors read and approved the final version of the manuscript.<br />

Acknowledgements<br />

We sincerely thank Mrs. Jette Lyby Michelsen and Mrs. Kristine J.K. Sørensen for<br />

technical assistance. Funding came from the 2008 Faculty of Science, University<br />

of Copenhagen Freja-Programme and from the Carlsberg Foundation<br />

(grant numbers: 2004_04_0572; 2006_01_0534; 2008_01_0466). The funders<br />

had no role in study design, data collection and analysis, decision to publish, or<br />

preparation of the manuscript.<br />

Author Details<br />

Department of Biology, University of Copenhagen, Universitetsparken, DK-<br />

2100 Copenhagen, Denmark<br />

Received: 18 September 2009 Accepted: 27 May 2010<br />

Published: 27 May 2010<br />

© This BMC 2010 is article Developmental an Haugan Open is available Access et al; Biology licensee from: article 2010, http://www.biomedcentral.com/1471-213X/10/56<br />

distributed BioMed 10:56 Central under Ltd. the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.<br />

References<br />

1. Evans DH: Osmotic and ionic regulation: Cells and animals Boca Raton: CRC<br />

Press; 2009.<br />

2. Vize PD, Woolf AS, Bard JBL: The Kidney: From normal development to<br />

congenital disease London: Academic Press; 2003.<br />

3. Saxén L: Organogenesis of the kidney Cambridge: Cambridge University<br />

Press; 1987.<br />

4. Dantzler WH: Comparative aspects of renal function. In The Kidney:<br />

Physiology and Pathophysiology 2nd edition. Edited by: Seldin D, Giebisch<br />

G. New York: Raven Press; 1992:885-942.<br />

5. Møbjerg N, Jespersen A, Wilkinson M: Morphology of the kidney in the<br />

West African caecilian, Geotrypetes seraphini (Amphibia, Gymnophiona,<br />

Caeciliidae). Journal of Morphology 2004, 262:583-607.<br />

6. Kriz W, Kaissling B: Structural and functional organization of the<br />

mammalian kidney. In Seldin and Giebisch's the kidney. Physiology and<br />

pathophysiology Edited by: Alpern RJ, Herbert SC. Burlington, MA:<br />

Academic Press; 2008:479-563.<br />

7. Greger R: Principles of renal transport; concentration and dilution of<br />

urine. In Comprehensive human physiology. From cellular mechanisms to<br />

integration Volume 2. Edited by: Greger R, Windhorst U. Berlin Heidelberg:<br />

Springer; 1996:1489-1516.<br />

8. Giebisch G, Windhager E: Urine concentration and dilution. In Medical<br />

physiology. A cellular and molecular approach Edited by: Boron WF,<br />

Boulpaep EL. Philadelphia: Saunders; 2005:828-844.<br />

9. Larsen EH, Møbjerg N, Nielsen R: Application of the Na + recirculation<br />

theory to ion coupled water transport in low- and high resistance<br />

osmoregulatory epithelia. Comparative Biochemistry and Physiology<br />

2007, 148:101-116.<br />

10. Larsen EH, Møbjerg N: Na + recirculation and isosmotic transport.<br />

Journal of Membrane Biology 2006, 212:1-15.<br />

11. Brandli AW: Towards a molecular anatomy of the Xenopus pronephric<br />

kidney. International Journal of Developmental Biology 1999, 43:381-395.<br />

12. Drummond IA, Majumdar A, Hentschel H, Elger M, Solnica-Krezel L, Schier<br />

AF, Neuhauss SCF, Stemple DL, Zwartkruis F, Rangini Z, et al.: Early<br />

development of the zebrafish pronephros and analysis of mutations<br />

affecting pronephric function. Development 1998, 125:4655-4667.<br />

13. Hillyard SD, Møbjerg N, Tanaka S, Larsen EH: Osmotic and ion regulation<br />

in amphibians. In Osmotic and ionic regulation: Cells and animals 1st<br />

edition. Edited by: Evans DH. Boca Raton: CRC Press; 2009:367-441.<br />

14. Møbjerg N, Larsen EH, Jespersen A: Morphology of the kidney in larvae<br />

of Bufo viridis (Amphibia, Anura, Bufonidae). Journal of Morphology<br />

2000, 245:177-195.<br />

15. Vize PD: Embryonic kidneys and other nephrogenic models. In The<br />

Kidney: From normal development to congenital disease Edited by: Vize PD,<br />

Woolf AS, Bard JBL. London: Academic Press; 2003:1-6.<br />

16. Vize PD, Seufert DW, Carroll TJ, Wallingford JB: Model systems for the<br />

study of kidney development: Use of the pronephros in the analysis of<br />

organ induction and patterning. Developmental Biology 1997,<br />

188:189-204.<br />

17. Zhou XL, Vize PD: Proximo-distal specialization of epithelial transport<br />

processes within the Xenopus pronephric kidney tubules.<br />

Developmental Biology 2004, 271:322-338.<br />

18. Tran U, Pickney LM, Ozpolat BD, Wessely O: Xenopus Bicaudal-C is<br />

required for the differentiation of the amphibian pronephros.<br />

Developmental Biology 2007, 307:152-164.<br />

19. Raciti D, Reggiani L, Geffers L, Jiang Q, Bacchion F, Subrizi AE, Clements D,<br />

Tindal C, Davidson DR, Kaissling B, et al.: Organization of the pronephric<br />

kidney revealed by large-scale gene expression mapping. Genome<br />

Biology 2008, 9:.<br />

20. Wingert RA, Davidson AJ: The zebrafish pronephros: A model to study<br />

nephron segmentation. Kidney International 2008, 73:1120-1127.<br />

21. Vasilyev A, Liu Y, Mudumana S, Mangos S, Lam PY, Majumdar A, Zhao JH,<br />

Poon KL, Kondrychyn I, Korzh V, et al.: Collective cell migration drives<br />

morphogenesis of the kidney nephron. Plos Biology 2009, 7:101-114.<br />

22. Wheeler GN, Brandli AW: Simple vertebrate models for chemical<br />

genetics and drug discovery screens: Lessons from Zebrafish and<br />

Xenopus. Developmental Dynamics 2009, 238:1287-1308.<br />

23. Vize PD, Carroll TJ, Wallingford JB: Induction, development and<br />

physiology of the pronephric tubules. In The Kidney: From normal<br />

development to congenital disease Edited by: Vize PD, Woolf AS, Bard JBL.<br />

London: Academic Press; 2003:19-50.<br />

24. Christensen AK: Structure of functional pronephros in larvae of<br />

Ambystoma opacum as studied by light and electron microscopy.<br />

American Journal of Anatomy 1964, 115:257-278.<br />

25. Eid SR, Brandli AW: Xenopus Na, K-ATPase: primary sequence of the beta<br />

2 subunit and in situ localization of alpha 1, beta 1, and gamma<br />

expression during pronephric kidney development. Differentiation<br />

2001, 68:115-125.<br />

26. Kumano T, Konno N, Wakasugi T, Matsuda K, Yoshizawa H, Uchiyama M:<br />

Cellular localization of a putative Na + /H + exchanger 3 during ontogeny<br />

in the pronephros and mesonephros of the Japanese black<br />

salamander (Hynobius nigrescens Stejneger). Cell and Tissue Research<br />

2008, 331:675-685.<br />

27. Uochi T, Takahashi S, Ninomiya H, Fukui A, Asashima M: The Na + , K + -<br />

ATPase alpha subunit requires gastrulation in the Xenopus embryo.<br />

Development Growth & Differentiation 1997, 39:571-580.<br />

28. Field HH: The development of the pronephros and segmental duct in<br />

Amphibia Cambridge, USA: Bulletin of the Museum of Comparative<br />

Zoology; 1891.<br />

29. Howland RB: On the effect of removal of the pronephros of the<br />

amphibian embryo. Proceedings of the National Academy of Sciences of<br />

the United States of America 1916, 2:231-234.<br />

30. Drawbridge J, Steinberg MS: Morphogenesis of the axolotl pronephric<br />

duct: A model system for the study of cell migration in vivo.<br />

International Journal of Developmental Biology 1996, 40:709-713.<br />

31. Gillespie LL, Armstrong JB: Formation of the pronephros and pronephric<br />

duct rudiment in the Mexican axolotl. Journal of Morphology 1985,<br />

185:217-222.<br />

32. Poole TJ, Steinberg MS: Amphibian pronephric duct morphogenesis -<br />

segregation, cell rearrangement and directed migration of the<br />

Ambystoma duct rudiment. Journal of Embryology and Experimental<br />

Morphology 1981, 63:1-16.<br />

33. Jaffee OC: Morphogenesis of the pronephros of the leopard frog (Rana<br />

pipiens). Journal of Morphology 1954, 95:109-123.<br />

34. Dietl P, Stanton BA: The amphibian distal nephron. In New insights in<br />

vertebrate kidney function Edited by: Brown JA, Balment RJ, Rankin JC.<br />

Cambridge: Cambridge University Press; 1993:115-134.<br />

35. Horisberger J, Hunter M, Stanton B, Giebisch G: The collecting tubule of<br />

Amphiuma II. Effects of potassium adaptation. American Journal of<br />

Physiology 1987, 253:F1273-F1282.<br />

36. Hunter M, Horisberger J-D, Stanton B, Giebisch G: The collecting tubule<br />

of Amphiuma I. Electrophysiological characterization. American Journal<br />

of Physiology 1987, 253:F1263-F1272.


Haugan et al. BMC Developmental Biology 2010, 10:56<br />

http://www.biomedcentral.com/1471-213X/10/56<br />

Page 9 of 9<br />

37. Møbjerg N, Larsen EH, Novak I: K + transport in the mesonephric<br />

collecting duct system of the toad Bufo bufo: microelectrode<br />

recordings from isolated and perfused tubules. Journal of Experimental<br />

Biology 2002, 205:897-904.<br />

38. Møbjerg N, Larsen EH, Novak I: Ion transport mechanisms in the<br />

mesonephric collecting duct system of the toad Bufo bufo:<br />

microelectrode recordings from isolated and perfused tubules.<br />

Comparative Biochemistry and Physiology A-Molecular & Integrative<br />

Physiology 2004, 137:585-595.<br />

39. Møbjerg N, Werner A, Hansen SM, Novak I: Physiological and molecular<br />

mechanisms of inorganic phosphate handling in the toad Bufo bufo.<br />

Pflugers Archiv - European Journal of Physiology 2007, 454:101-113.<br />

40. Koeppen BM: Conductive properties of the rabbit outer medullary<br />

collecting duct: outer stripe. American Journal of Physiology - Renal<br />

Physiology 1986, 250:F70-F76.<br />

41. Schlatter E: Regulation of ion channels in the cortical collecting duct.<br />

Renal Physiology and Biochemistry 1993, 16:21-36.<br />

42. Wang WH, Giebisch G: Regulation of potassium (K) handling in the renal<br />

collecting duct. Pflugers Archiv-European Journal of Physiology 2009,<br />

458:157-168.<br />

43. Garty H, Palmer LG: Epithelial sodium channels: Function, structure, and<br />

regulation. Physiological Reviews 1997, 77:359-396.<br />

44. Loffing J, Korbmacher C: Regulated sodium transport in the renal<br />

connecting tubule (CNT) via the epithelial sodium channel (ENaC).<br />

Pflugers Archiv-European Journal of Physiology 2009, 458:111-135.<br />

45. Schlatter E, Greger R, Schafer JA: Principal cells of cortical collecting<br />

ducts of the rat are not a route of transepithelial Cl - transport. Pflugers<br />

Archiv-European Journal of Physiology 1990, 417:317-323.<br />

46. Bordzilovskaya NP, Dettlaf TA, Duhan ST, Malacinski GM: Developmentalstage<br />

series of axolotl embryos. In Developmental Biology of the Axolotl<br />

Edited by: Armstrong JB, Malacinski GM. New York: Oxford University<br />

Press; 1989:201-219.<br />

47. Greger R, Hampel W: A modified system for in vitro perfusion of isolated<br />

renal tubules. Pflugers Archiv - European Journal of Physiology 1981,<br />

389:175-176.<br />

doi: 10.1186/1471-213X-10-56<br />

Cite this article as: Haugan et al., Functional characterization of the vertebrate<br />

primary ureter: Structure and ion transport mechanisms of the pronephric<br />

duct in axolotl larvae (Amphibia) BMC Developmental Biology 2010,<br />

10:56


Paper V


2803<br />

The Journal of Experimental Biology 212, 2803-2811<br />

Published by The Company of Biologists 2009<br />

doi:10.1242/jeb.029413<br />

Cyclomorphosis in Tardigrada: adaptation to environmental constraints<br />

<strong>Kenneth</strong> Agerlin <strong>Halberg</strong> 1 , Dennis Persson 1,2 , Hans Ramløv 3 , Peter Westh 3 , Reinhardt Møbjerg Kristensen 2<br />

and Nadja Møbjerg 1, *<br />

1 Department of Biology, University of Copenhagen, August Krogh Building, Universitetsparken 13, DK-2100 Copenhagen Ø,<br />

Denmark, 2 Natural History Museum of Denmark, Zoological Museum, Invertebrate Department, Universitetsparken 15, DK-2100<br />

Copenhagen Ø, Denmark and 3 Department of Nature, Systems and Models, University of Roskilde, Universitetsvej 1, DK-4000<br />

Roskilde, Denmark<br />

*Author for correspondence (e-mail: nmobjerg@bio.ku.dk)<br />

Accepted 9 June 2009<br />

SUMMARY<br />

Tardigrades exhibit a remarkable resilience against environmental extremes. In the present study, we investigate mechanisms of<br />

survival and physiological adaptations associated with sub-zero temperatures and severe osmotic stress in two commonly found<br />

cyclomorphic stages of the marine eutardigrade Halobiotus crispae. Our results show that only animals in the so-called<br />

pseudosimplex 1 stage are freeze tolerant. In pseudosimplex 1, as well as active-stage animals kept at a salinity of 20 ppt, ice<br />

formation proceeds rapidly at a crystallization temperature of around –20°C, revealing extensive supercooling in both stages,<br />

while excluding the presence of physiologically relevant ice-nucleating agents. Experiments on osmotic stress tolerance show<br />

that the active stage tolerates the largest range of salinities. Changes in body volume and hemolymph osmolality of active-stage<br />

specimens (350–500 μm) were measured following salinity transfers from 20 ppt. Hemolymph osmolality at 20 ppt was<br />

approximately 950 mOsm kg –1 . Exposure to hypo-osmotic stress in 2 and 10 ppt caused (1) rapid swelling followed by a regulatory<br />

volume decrease, with body volume reaching control levels after 48 h and (2) decrease in hemolymph osmolality followed by a<br />

stabilization at significantly lower osmolalities. Exposure to hyperosmotic stress in 40 ppt caused (1) rapid volume reduction,<br />

followed by a regulatory increase, but with a new steady-state after 24 h below control values and (2) significant increase in<br />

hemolymph osmolality. At any investigated external salinity, active-stage H. crispae hyper-regulate, indicating a high water<br />

turnover and excretion of dilute urine. This is likely a general feature of eutardigrades.<br />

Key words: cyclomorphosis, environmental stress, freeze tolerance, Halobiotus crispae, invertebrate, osmoregulation, tardigrade, volume<br />

regulation.<br />

INTRODUCTION<br />

The phylum Tardigrada comprises a group of hydrophilous micrometazoans,<br />

exhibiting close affinities to the euarthropod complex<br />

(Garey et al., 1996; Giribet et al., 1996; Mallatt et al., 2004). They<br />

occupy a range of niches in terrestrial, freshwater and marine<br />

environments from continental Antarctica (Convey and McInnes,<br />

2005) to the icecap of Greenland (Grøngaard et al., 1999) yet are<br />

especially abundant in mosses and lichens, where they constitute a<br />

major component of the cryptic fauna. Along with nematodes and<br />

rotifers, selected species of tardigrades exhibit a remarkable<br />

resilience against physical extremes, including low and high<br />

temperatures (–253°C to +151°C), ionizing radiation (up to<br />

6000Gy), vacuum, high pressure (up to 600MPa) and extreme<br />

desiccation (Ramløv and Westh, 1992; Westh and Kristensen, 1992;<br />

Ramløv and Westh, 2001; Schill et al., 2004; Horikawa et al., 2006;<br />

Jönson and Schill, 2007; Hengherr et al., 2008; Hengherr et al.,<br />

2009). However, the underlying physiological and biochemical<br />

mechanisms mediating these unique tolerances are still largely<br />

unidentified and represent an exciting challenge to contemporary<br />

biology.<br />

The marine eutardigrade Halobiotus crispae Kristensen 1982<br />

(Fig.1) colonizes tidal and subtidal habitats at numerous localities<br />

throughout the northern hemisphere (Møbjerg et al., 2007). This<br />

species is characterized by the appearance of seasonal cyclic<br />

changes in morphology, i.e. cyclomorphosis (Kristensen, 1982).<br />

Three distinct cyclomorphic stages have been recognized: (1) the<br />

active stage, (2) the pseudosimplex 1 (P1) stage and (3) the<br />

pseudosimplex 2 (P2) stage (Møbjerg et al., 2007). The defining<br />

physiological and biochemical characteristics of the individual stages<br />

are largely unknown but most likely correlate with dominant abiotic<br />

factors. A ubiquitous factor in all tidal and subtidal habitats is the<br />

large temporal and spatial fluctuations in external salinity. Yet,<br />

additional adaptations are necessary at high latitudes to ensure winter<br />

survival due to prolonged exposure to subzero temperatures. In the<br />

present study, we focus on the adaptive significance of the two main<br />

cyclomorphic stages in H. crispae, i.e. the active stage corresponding<br />

to the reproductive stage of other tardigrades and the P1 stage, a<br />

hibernation stage, which is comparable to the cysts found in other<br />

tardigrades (e.g. Guidetti et al., 2008). We do not deal with the P2<br />

stage, which is a sexual maturation stage that has not yet been<br />

reported from other tardigrades. Our preliminary data, however,<br />

suggest that this stage has a unique osmoregulatory profile. We show<br />

that the transition between the active and P1 stages is associated<br />

with profound changes in the physiology of the animal. The P1 stage<br />

is the only stage at which H. crispae survives internal ice formation.<br />

The active stage tolerates large shifts in ambient salinity and we<br />

investigate in detail the volume and osmoregulatory capacity of this<br />

stage. Our study presents the first detailed analysis of osmoregulation<br />

in tardigrades. The data show that active-stage H. crispae hyperregulate<br />

at any investigated external salinity, which would indicate<br />

excretion of dilute urine. This is likely to be a general feature of<br />

eutardigrades, which all possess Malpighian tubules.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


2804<br />

K. A. <strong>Halberg</strong> and others<br />

Fig. 1. SEM investigation of Halobiotus crispae from Vellerup Vig, Denmark. (A) Overview of P1 stage indicating the areas shown in B and C. The thick outer<br />

cuticle functionally isolates the animal from the surroundings (scale bar=100 μm). (B) Close-up of the head region of P1. Notice that the mouth is closed by<br />

cuticular thickenings (scale bar=25 μm). (C) Close-up of the posterior area of the P1 stage. As shown for the mouth, the cloaca is closed (scale bar=10 μm).<br />

(D) Close-up of the head region of the active stage. Note the six peribuccal sensory organs (*) that surround the open mouth (scale bar=25 μm). (E) Close-up<br />

of the posterior area of the active stage, revealing the open tri-lobed cloaca (scale bar=10 μm).<br />

MATERIALS AND METHODS<br />

Tardigrade sampling<br />

Specimens of Halobiotus crispae were collected at regular intervals<br />

in the period 2005 to 2008 at Vellerup Vig, Isefjord, Denmark<br />

(55°44.209N, 11°51.365E) and at Nipisat Bay, Disko Island, West<br />

Greenland (69°25.934N, 54°10.768E) in August 2006. At Vellerup<br />

Vig, bottom samples were collected at an approximate depth of<br />

1.0–2.5m, while samples from Nipisat Bay were taken in the subtidal<br />

zone 3–4cm below low tide. With the exception of the data on<br />

osmotic stress tolerance presented in Fig.3, the obtained results are<br />

based entirely on animals collected at Vellerup Vig. Detailed<br />

descriptions of the two localities can be found elsewhere (Kristensen,<br />

1982; Møbjerg et al., 2007). Collected samples were freshwatershocked,<br />

decanted into a conical net (mesh size 62 μm) and<br />

transferred to Petri dishes. These dishes were supplied with fresh<br />

seawater (SW; 18–20ppt; pH8–9) and substrate from the locality.<br />

Tardigrades were localized using a Leica MZ16 stereomicroscope<br />

(Leica Microsystems, Wetzlar, Germany), and primarily found on<br />

the haptera of various filamentous algae present in the substrate.<br />

Isolated tardigrades from Vellerup Vig were kept at 4°C in SW for<br />

periods of up to 6months by regularly supplying fresh substrate<br />

from the locality. Different cyclomorphic stages (see Fig.1) were<br />

identified using an Olympus BX 51 interference-contrast microscope<br />

(Olympus, Tokyo, Japan).<br />

Scanning electron microscopy<br />

For scanning electron microscopy, specimens were fixed in 2.5%<br />

glutaraldehyde in 0.1moll –1 sodium cacodylate buffer (pH7.4),<br />

rinsed in the buffer and subsequently postfixed in 1% OsO 4 in<br />

0.1moll –1 sodium cacodylate buffer (pH7.4). Following fixation,<br />

the specimens were dehydrated through a graded series of ethanol<br />

and acetone. They were critical point dried (Bal-Tec CPD 030 critical<br />

point dryer, Bal-Tec Union, Balzers, Liechtenstein), mounted on<br />

aluminum stubs, sputter-coated with platinum–palladium (thickness<br />

~12nm) using a JEOL JFC-2300HR (JEOL, Tokyo, Japan) and<br />

examined in a JEOL JSM-6335F Field Emission scanning electron<br />

microscope (JEOL, Japan).<br />

Cold hardiness<br />

Six groups of 10 animals in both the active and P1 stage were<br />

transferred to Eppendorf tubes containing 1.5ml of SW from<br />

Vellerup Vig (20ppt). The samples were cooled to a constant<br />

temperature of –20°C at a cooling rate of approximately 1°Cmin –1<br />

(Block, 1991) and held at the target temperature for a period of 24h.<br />

The animals were thawed at room temperature and the survival<br />

assessed successively over the course of 96h. Animals retaining<br />

locomotory function or responsive to tactile stimuli following this<br />

period were considered alive.<br />

As subzero temperatures may be experienced for longer periods<br />

of time in Arctic habitats, the long-term survival at subzero<br />

temperatures was investigated. An additional six groups of 10<br />

specimens in each cyclomorphic stage were frozen to –20°C at a<br />

cooling rate of approximately 1°Cmin –1 and kept frozen for a total<br />

of 36days. Survival was assessed as described above.<br />

Differential scanning calorimetry<br />

The quantity and kinetics of ice formation associated with cooling<br />

of H. crispae from Vellerup Vig (20ppt) in active and P1 stages<br />

were studied by differential scanning calorimetry (DSC). Groups<br />

of 40–75 animals in each respective stage were transferred to 30μl<br />

aluminum DSC pans. In order to avoid dehydrating the animals<br />

during the removal of external water, the tardigrades were clumped<br />

in the central part of the DSC pan and excess water was subsequently<br />

removed with small pieces of delicate task wipes. Sample mass was<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Cyclomorphosis in tardigrades<br />

2805<br />

determined gravimetrically to the nearest 0.01mg using a fine-scale<br />

AT261 Deltarange (Mettler-Toledo, Columbus, OH, USA), yielding<br />

a total mass of 0.16–0.39mg (wet mass). The pans were sealed and<br />

transferred to a calorimeter (Perkin Elmer DSC 7 equipped with an<br />

Intercooler II mechanical cooling device), with an empty pan as<br />

reference. The calorimeter was calibrated with gallium [melting<br />

point, T m =29.78°C; melting enthalpy, H m =80.1 J g –1 ], water<br />

(T m =0°C) and n-decane (T m =–29.66°C). All scans involved cooling<br />

from 5°C to –40°C and subsequent reheating to 5°C at a cooling<br />

rate of 5°Cmin –1 . Samples were reweighed following the DSC run<br />

to ensure that no water loss had occurred. In order to determine the<br />

water content of samples following the freeze/thaw cycle, pans were<br />

punctured and dried at 80°C to a constant mass (dry mass). A<br />

minimum of three groups of animals in each stage was used (see<br />

Table1). The obtained thermograms (heat flows vs temperature)<br />

were analyzed with respect to crystallization temperature (T c ),<br />

amount of ice formed during the freezing exotherm (assuming that<br />

the latent heat of crystallization is the same as for pure water), T m<br />

and the osmotic pressure of the extracellular fluids as calculated by<br />

the standard DSC 7 software. Ice contents were calculated using<br />

the water content and the enthalpy of the freeze exotherm. The<br />

temperature dependence of the enthalpy of crystallization of water<br />

was taken into account as previously described (Kristiansen and<br />

Westh, 1991). For hemolymph osmolality calculations, the onset<br />

melting point measured by the DSC 7 software was determined<br />

according to the approach of Nicholajsen and Hvidt (Nicholajsen<br />

and Hvidt, 1994), in which the established melting point of the body<br />

fluid was derived from a standard curve made from predetermined<br />

NaCl solutions.<br />

Preparation of experimental solutions<br />

Salt water solutions of different osmotic pressure were made by<br />

successive dilution with distilled water or by evaporative reduction<br />

of 100% SW from the locality. Measurements of osmotic pressure<br />

were made in parallel on a Vapro 5520 vapor pressure osmometer<br />

(Wescor, Logan, UT, USA) and on a refractometer (S-1 Shibuya<br />

Land, Tokyo, Japan).<br />

Osmotic stress tolerance<br />

Animals collected from the Danish as well as the Greenlandic<br />

population were used for the current experiment. Groups of 20<br />

specimens were transferred to small glass vials containing 4ml of<br />

100% SW at 4°C, and specimens were either exposed to a gradual<br />

increase or decrease in salinity. The gradual changes in salinity were<br />

performed over the course of 4–5h by periodically replacing small<br />

volumes of SW with prefixed solutions of either a higher or lower<br />

salinity. Animal activity was concomitantly assessed. Animals were<br />

allowed a period of 20–40min of acclimatization following a salinity<br />

change prior to assessment. Individuals responsive to tactile stimuli<br />

were considered active. Five groups of specimens in active and P1<br />

stages were assessed at both hypo- and hyperosmotic salinities.<br />

Volume measurements<br />

Individual adult active-stage specimens of H. crispae (size<br />

300–500μm) from Vellerup Vig (20ppt) were visualized in an<br />

Olympus BX 51 microscope (Olympus), photographed using a<br />

digital camera (C-5050, Olympus) and subsequently exposed for<br />

set time periods of 30min, 1, 2, 4, 24 and 48h to saltwater solutions<br />

with salinities of 2ppt, 10ppt and 40ppt. The osmotic treatments<br />

were conducted in small glass vials containing 4ml of SW at 4°C.<br />

At the end of each time interval, individuals were transferred, in a<br />

drop of the appropriate solution, to glass microscope slides and<br />

photographed under cover slips for subsequent estimations of body<br />

volume. During photography, great care was taken to minimize the<br />

time spent by the animals under the cover slips, in order to avoid<br />

evaporative water loss, which would alter the osmotic pressure of<br />

the solution. The animals were ensured total freedom of movement.<br />

Following photography, individuals were returned to the respective<br />

salinities until the end of the next set time period, when the process<br />

was repeated. At each of the time intervals, 10–14 individuals were<br />

photographed at each of the SW treatments. Images were analyzed<br />

using DP-soft TM (Olympus), and total body volume was calculated<br />

according to the equation: V total =π(r 2 bodyh body +2r 2 legh leg ), where V<br />

is the volume of the specimen, r is the measured radius, and h is<br />

the measured length of the body and hind legs, respectively.<br />

In order to assess the behavioral response of H. crispae during<br />

osmotic shock and to quantify potential mortality related to the<br />

respective treatments, a separate experiment was performed.<br />

Specimens (N=10) were transferred directly to 2ppt, 10ppt, 20ppt<br />

(control) and 40 ppt, respectively, and animal activity was<br />

subsequently monitored over the course of 48h at 4°C. Individuals<br />

responsive to tactile stimuli at the above-mentioned set time periods<br />

were considered active and alive (see Fig.7). Three groups exposed<br />

to each treatment were assessed.<br />

Measurement of hemolymph osmolality<br />

Hemolymph osmolality was measured in individual tardigrades<br />

following exposure for 30min, 4 and 48h to the experimental<br />

solutions of 2ppt (62mOsmkg –1 ), 10ppt (311mOsmkg –1 ) and 40ppt<br />

(1245 mOsm kg –1 ). Six animals were used for osmolality<br />

determination in each of the experimental solutions. Six animals<br />

kept at 20ppt (623mOsmkg –1 ) served as a control. Hemolymph<br />

Table 1. Post-freeze survival and data obtained from differential scanning calorimetry on H. crispae from Vellerup Vig, Denmark (20 ppt)<br />

in the active and P1 stages<br />

Sample<br />

Post-freeze<br />

survival (%)<br />

(frozen for 24 h)<br />

Post-freeze<br />

survival (%)<br />

(frozen for<br />

36 days)<br />

Crystallization<br />

temp. (°C)<br />

Melting temp.<br />

(°C)<br />

Water content<br />

(%)<br />

Body-water<br />

frozen during<br />

freezing<br />

exotherm (%)<br />

Osmolality of<br />

extracellular<br />

–1<br />

fluids (mOsm kg )<br />

Pseudosimplex 1 53.3±15 (6) 12.7±7 (6) –19.6±3.1 (6) –4.29±0.79 (5) 67±4 (5) 59±3 (4) 928±77 (5)<br />

Active 0 (6) 0 (6) –21.6±2.1 (3) –4.48±0.15 (3) 68±4 (3) 69±5 (3) 975±36 (3)<br />

t-test (P0.05 (NS, not significant), P


2806<br />

K. A. <strong>Halberg</strong> and others<br />

samples (2–3nl) were collected by piercing individual specimens<br />

under immersion oil (type A; 150 centistoke; Cargille Laboratories,<br />

Cedar Grove, NJ, USA) using hand-pulled glass capillary tubes<br />

(capacity 1 μl; Micro-caps, Drummond Scientific Company,<br />

Broomall, PA, USA). Hemolymph samples were acquired through<br />

capillary action and subsequently ejected into immersion oil. Care<br />

was taken to ensure that the measurements were made on fluid<br />

originating from hemolymph alone and samples containing gut<br />

contents were discarded. Prior to sample collection, immersion oil<br />

was collected into the capillary tube in order to avoid any evaporative<br />

water loss. Using an Irvin loop, samples were immediately<br />

transferred in a drop of immersion oil into sample oil wells (type<br />

B; 1250 centistoke; Cargille Laboratories) of a calibrated nanoliter<br />

osmometer (Clifton Technical Physics, Hartford, NY, USA), and<br />

the osmolality (mOsmkg –1 ) was determined by melting point<br />

depression (MDP=1.858°COsm –1 ).<br />

Statistics<br />

Significant differences between experimental and control conditions<br />

were tested using unpaired, two-tailed t-tests, and a significance level<br />

of P≤0.05.<br />

RESULTS<br />

Survival at sub-zero temperatures<br />

The external morphology of H. crispae in the active and the P1<br />

stage is shown in Fig.1. Notably, the P1 stage is characterized by<br />

a conspicuous double cuticle, in which both the mouth and cloaca<br />

are closed by cuticular thickenings (Fig.1A–C). Post-freeze survival<br />

following both short- and long-term exposure to subzero<br />

temperatures is listed in Table1. An example of the quantitative<br />

kinetics of ice formation associated with the freeze/thaw cycle of<br />

P1-stage H. crispae kept at 20ppt is illustrated in Fig.2. The<br />

collective DSC analysis of the onsets and areas of peaks after the<br />

freeze/thaw cycles, together with the water content of the samples,<br />

provide the remaining results listed in Table1.<br />

Whereas animals in the active stage and P2 stage (data not shown)<br />

were intolerant of freezing, animals in the P1 stage were<br />

demonstrated to be freeze-tolerant. Post-freeze survival following<br />

24h exposure to –20°C was 53.3±15%; however, when prolonging<br />

the period spent at subzero temperatures to 36days, the survival of<br />

animals in the P1 stage decreased to 12.7±7% (Table1). Animal<br />

recovery was monitored over a period of 96 h following the<br />

freeze/thaw cycle; however, the majority of animals had resumed<br />

activity after a period of 48h. No additional recovery was monitored<br />

beyond 96h following any of the investigated treatments.<br />

Substantial ice formation proceeded rapidly following the first<br />

ice nucleation in the freeze-tolerant P1 stage with a mean of<br />

–19.6±3.1°C; as indicated by the large exotherm in Fig.2. At the<br />

given salinity, the amount of ice formed during the freezing<br />

exotherm amounted to approximately 60% of the body water<br />

(Table1). The absence of additional small exothermic peaks during<br />

the subsequent cooling to –40°C indicated that no additional ice<br />

formation occurred following the initial large freezing exotherm.<br />

The freeze exotherm lasted less than one minute. The initial<br />

separation of ice (T c ) occurred in the temperature range of –15.4 to<br />

–23.2°C.<br />

Only marginal differences in T c of animals in the P1 and active<br />

stage were observed (Table1), suggesting an absence of seasonal<br />

variations in ice-nucleating activity in H. crispae. In spite of invariant<br />

water contents between the two stages, the amount of water<br />

crystallizing during cooling in animals in the active stage was<br />

significantly higher than in animals in the P1 stage; however, the<br />

melting points of the two stages remained largely unaltered. The<br />

latter indicates that body fluid osmolality at a given external salinity<br />

is unaffected by the animal’s transition from the active to the P1<br />

stage.<br />

Volume- and osmoregulatory capacity<br />

Fig.3 shows the percentage of active animals of H. crispae following<br />

the exposure to gradual changes in the external salinity. When<br />

comparing Danish P1 and active-stage H. crispae, the active stage<br />

displayed a larger tolerance towards the more concentrated SW<br />

solutions and were slightly more tolerant of the very dilute solutions.<br />

Indeed, a significantly higher percentage of active-stage specimens<br />

Differential heat flow (mW)<br />

15<br />

7.5<br />

Exotherm<br />

Cooling<br />

Heating<br />

Endotherm<br />

0<br />

–40 –35 –30 –25 –20 –15 –10 –5 0 5<br />

Temperature (°C)<br />

Fig. 2. Representative thermogram displaying the exothermic and<br />

endothermic events associated with the cooling and heating of a sample<br />

containing P1-stage Halobiotus crispae from Vellerup Vig, Denmark<br />

(20 ppt). The crystallization and melting temperatures were estimated as<br />

the onsets of peaks (–17.5°C and –4.5°C, respectively) using the DSC 7<br />

software (see also Table 1). The very low crystallization temperatures<br />

measured exclude the presence of physiologically relevant ice-nucleating<br />

agents.<br />

Activity (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Active, Vellerup<br />

P1, Vellerup<br />

Active, Nipisat<br />

Osmolality (mOsm kg –1 )<br />

0 500 1000 1500 2000 2500<br />

*<br />

*<br />

0 10 20 30 40 50 60 70 80 90<br />

Salinity (ppt)<br />

100<br />

Fig. 3. Osmotic stress tolerance of Halobiotus crispae in the P1 and active<br />

stages. Active-stage () and P1-stage () specimens from the Danish<br />

population at Vellerup Vig. Active-stage () specimens from the<br />

Greenlandic population at Nipisat Bay. Data are means ± s.e.m. from five<br />

independent experiments. *, significantly different from Vellerup Vig P1<br />

stage (P


Cyclomorphosis in tardigrades<br />

2807<br />

remained active in the salinity spectrum of 0–3ppt and 45–60ppt,<br />

as compared with the animals in the P1 stage, and in general seemed<br />

less affected by the impositions of osmotic stress. Our preliminary<br />

data on P2 from Vellerup Vig show that this stage tolerates very<br />

dilute solutions better than the other stages, yet is the least tolerant<br />

of increases in salinity, becoming inactive at around 50ppt. As a<br />

comparison, the active stage from Greenland (Nipisat Bay), living<br />

in a more exposed habitat compared with the Vellerup Vig<br />

population, displayed an even higher tolerance to concentrated SW,<br />

with observed activity at 80ppt.<br />

In the following, we investigate in detail volume and<br />

osmoregulation in active-stage H. crispae from Vellerup Vig kept<br />

at a control salinity of 20ppt (Figs4–8). When exposed to a severe<br />

hypo-osmotic shock of 2 ppt (63 mOsm kg –1 ), the animals<br />

exhibited a large significant increase in body volume, resulting<br />

in a total body volume of 127±11% after merely 0.5h of exposure<br />

(Fig.4A,B). During this time period, animals became bloated and<br />

rigid and most specimens lost locomotory functions (Fig.7A).<br />

This passive uptake of water continued during the initial 2h of<br />

the exposure, culminating in a total body volume of 162±17%.<br />

However, after this period of time, a regulatory volume decrease<br />

(RVD) was observed. The total body volume of specimens was<br />

considerably reduced to 114±14% following 48h immersion and<br />

was not significantly different from the controls (Fig. 4B).<br />

Additionally, an increase in the number of active animals was<br />

observed, yet a considerable number of specimens remained<br />

passive throughout the treatment (Fig. 7A). Nevertheless,<br />

following a gradual return to 20 ppt, all animals regained<br />

locomotory functions. No mortality was observed in any of the<br />

treatments.<br />

Upon immersion of individual specimens into a less severe hypoosmotic<br />

media of 10ppt (311mOsmkg –1 ) a similarly significant<br />

increase in total body volume was observed; reaching a mean value<br />

of 132±11% after 0.5h exposure (Fig.5A,B). However, following<br />

this initial increase, total body volume stabilized, and a RVD was<br />

observed after 1–2h exposure. After 4h incubation, total body<br />

volume was 110±8%, which was not significantly different from<br />

the control situation. An effect on the locomotory functions was<br />

observed initially, as some animals displayed sluggish movements,<br />

yet only a limited number of animals were passive during this<br />

experiment (Fig.7B).<br />

When transferring H. crispae to a hyperosmotic solution of 40ppt<br />

(1245mOsmkg –1 ), a significant decrease in total body volume was<br />

observed (Fig.6A,B). Total body volume was significantly reduced<br />

to 66±9% following the first 0.5h of immersion and remained largely<br />

unaltered during the following hours of the treatment. After 24h,<br />

most specimens had displayed a regulatory volume increase,<br />

resulting in a mean total body volume of 82±9%, yet total body<br />

volume remained significantly different from the control situation<br />

even after 48h. Nevertheless, animal motility was little affected by<br />

the hypertonic shock, neither at the initial transfer nor throughout<br />

the rest of the experiment (Fig.7C).<br />

20 ppt,<br />

2 ppt,<br />

623 mOsm kg –1 62 mOsm kg –1<br />

A<br />

0h 0.5h 1h 2h 4h 24h 48h<br />

20 ppt,<br />

10 ppt,<br />

623 mOsm kg –1 311 mOsm kg –1<br />

A<br />

0h 0.5h 1h 2h 4h 24h 48h<br />

B<br />

Total body volume (%)<br />

200<br />

180<br />

160<br />

(12)<br />

(6)<br />

140<br />

(12)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

(6)<br />

(12)<br />

(12)<br />

(6)<br />

(10)<br />

0 1 2 3 4 20 30 40 50<br />

Exposure time (h)<br />

(10)<br />

(6)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

Hemolymph osmotic pressure (mOsm kg –1 )<br />

B<br />

Total body volume (%)<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

(6)(10)(12) (12)<br />

(10) (13)<br />

(6)<br />

(6)<br />

(11)<br />

0 1 2 3 4 20 30 40 50<br />

Exposure time (h)<br />

(6)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

Hemolymph osmotic pressure (mOsm kg –1 )<br />

Fig. 4. (A) Halobiotus crispae in active stage from Vellerup Vig, Denmark<br />

(20 ppt). Light-microscopical images at different time points following<br />

exposure to 2 ppt (62 mOsm kg –1 ) of a single specimen (scale bar=100 μm).<br />

(B) Changes in total body volume () and measured internal osmolality ()<br />

over a period of 48 h following exposure to an external salinity of 2 ppt.<br />

Data are expressed as means ± s.d. Numbers in parentheses indicate the<br />

number of animals used for assessment of body volume and hemolymph<br />

osmolality at each time point.<br />

Fig. 5. (A) Halobiotus crispae in active stage from Vellerup Vig, Denmark<br />

(20 ppt). Light-microscopical images at different time points following<br />

exposure to 10 ppt (311 mOsm kg –1 ) of a single specimen (scale<br />

bar=100 μm). (B) Changes in total body volume () and measured internal<br />

osmolality () over a period of 48 h following an exposure to an external<br />

salinity of 10 ppt. Data are expressed as means ± s.d. Numbers in<br />

parentheses indicate the number of animals used for assessment of body<br />

volume and hemolymph osmolality at each time point.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


2808<br />

K. A. <strong>Halberg</strong> and others<br />

20 ppt,<br />

40 ppt,<br />

623 mOsm kg –1 1245 mOsm kg –1<br />

A<br />

B<br />

Total body volume (%)<br />

0h 0.5h 1h 2h 4h 24h 48h<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

(6)<br />

(6)<br />

(12)<br />

(12) (14)<br />

(11)<br />

(13) (11)<br />

1400<br />

1200<br />

1000<br />

200<br />

40<br />

0 1 2 3 4 20 30 40 50<br />

Exposure time (h)<br />

Fig. 6 (A) Halobiotus crispae in active stage from Vellerup Vig, Denmark<br />

(20 ppt). Light-microscopical images at different time points following<br />

exposure to 40 ppt (1245 mOsm kg –1 ) of a single specimen (scale<br />

bar=100 μm). (B) Changes in total body volume () and measured internal<br />

osmolality () over a period of 48 h following an exposure to an external<br />

salinity of 40 ppt. Data are expressed as means ± s.d. Numbers in<br />

parentheses indicate the number of animals used for assessment of body<br />

volume and hemolymph osmolality at each time point.<br />

Notably, hemolymph osmotic pressure differed significantly<br />

from the control condition at all time points examined during the<br />

various salinity treatments (Figs4–6). Hemolymph osmolality of H.<br />

crispae varied in proportion to the gradient set up by the salinity<br />

transfer. The body fluids of animals kept under control conditions<br />

(20ppt, 623mOsmkg –1 ) had an osmolality of 926±29mOsmkg –1 ,<br />

(6)<br />

(6)<br />

800<br />

600<br />

400<br />

Hemolymph osmotic pressure (mOsm kg –1 )<br />

indicating that H. crispae hyper-regulates during steady-state<br />

conditions. This hyperosmotic regulation was independently<br />

confirmed by the DSC investigation in which the hemolymph<br />

osmolality was measured at 975±36mOsmkg –1 (Table1). The two<br />

measurements are not significantly different (t-test, P


Cyclomorphosis in tardigrades<br />

2809<br />

Hemolymph osmolality (mOsm kg –1 )<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

Active stage<br />

Osmotic performance<br />

Isoosmotic line<br />

0<br />

0 200 400 600 800 1000 1200 1400<br />

External osmolality (mOsm kg –1 )<br />

Fig. 8. Measured hemolymph osmolality of Halobiotus crispae (active stage<br />

from Vellerup Vig, Denmark) during steady-state conditions after 48 h<br />

acclimation to 2 ppt (63 mOsm kg –1 ), 10 ppt (311 mOsm kg –1 ), 20 ppt<br />

(623 mOsm kg –1 ) and 40 ppt (1245 mOsm kg –1 ), respectively. Each point<br />

represents the mean ± s.d. of the individual experiments. The broken line<br />

indicates the isoosmotic line at which no osmoregulation occurs.<br />

functionally characterized as a movable cyst (Kristensen, 1982;<br />

Møbjerg et al., 2007; Guidetti et al., 2008). In Denmark (Vellerup<br />

Vig), this stage is dominant during the summer months, presumably<br />

enabling H. crispae to withstand heat stress and oxygen depletion.<br />

The active stage, the only stage at which active feeding and sexual<br />

reproduction occur, is the dominant stage during the Greenlandic<br />

summer, whereas this stage is present during late winter and the<br />

spring months in Denmark (Møbjerg et al., 2007).<br />

Freeze avoidance and freeze tolerance<br />

Winter temperatures are frequently below the equilibrium freezing<br />

point of the surrounding seawater at least in some portions of the<br />

natural environments of H. crispae, and certain habitats may even<br />

become completely frozen for extended periods of the year<br />

(Kristensen, 1982). Enduring such hostile surroundings requires<br />

corresponding cold-tolerance strategies that enable long-term<br />

survival. Traditionally, two main options are exploited by<br />

ectothermic animals when faced with subzero temperatures, i.e.<br />

freeze avoidance and freeze tolerance (Lee, 1991). When exposing<br />

animals in the P1 stage of H. crispae to temperatures below the<br />

equilibrium freezing point (T c ) of their body fluids, freeze tolerance<br />

is demonstrated, indicating that winter survival could involve<br />

extracellular ice formation in this species. However, the finding that<br />

mortality increases with prolonged exposure to subzero temperatures<br />

suggests that the consequent damages accumulate in proportion to<br />

the time spent exposed to freezing conditions. This observation is<br />

likely explained by the depletion of essential metabolites and is of<br />

particular interest in Arctic habitats in which subfreezing<br />

temperatures have to be endured for long periods of time.<br />

In freeze-tolerant organisms, ice formation is usually promoted<br />

at relatively high subzero temperatures (–2 to –10°C) by icenucleating<br />

agents present in the extracellular fluid (Zachariassen,<br />

1985; Block, 1991; Westh and Kristensen, 1992). The adaptive<br />

advantage of such a strategy is that the process of ice formation<br />

proceeds relatively slowly at relatively high temperatures, enabling<br />

the organism to maintain the damage associated with freezing within<br />

tolerable boundaries. Indeed, both the localization and the amount<br />

of ice formed in freeze-tolerant organisms are usually under tight<br />

control. Surprisingly, our calorimetric investigation of the freezetolerant<br />

P1 stage reveals that ice crystallization occurs at<br />

approximately –20°C, excluding the presence of any physiologically<br />

relevant ice-nucleating agents in this stage. In fact, the very low<br />

crystallization temperatures measured in both stages suggest that<br />

the capacity for supercooling is maintained throughout the majority<br />

of the year, which is in general contrast to the pattern observed in<br />

most other freeze-tolerant invertebrates (Block, 1991; Westh and<br />

Kristensen, 1992; Ramløv et al., 1996). However, in spite of an<br />

invariant melting point and water content between the two stages,<br />

the amount of water crystallized during cooling to –40°C appears<br />

to be about 60% for animals in the P1 stage and 70% for animals<br />

in the active stage (see Table 1). Consequently, the cellular<br />

dehydration induced by freezing is expected to be much higher in<br />

animals in the active stage, compared with that of specimens in the<br />

P1 stage. This reduction in ice accumulation in specimens in the<br />

P1 stage could potentially be explained by an increased production<br />

of macromolecules [which kinetically inhibits ice formation but has<br />

negligible effect on the melting temperature (see Westh and<br />

Kristensen, 1992)], as compared with the active stage. Whether this<br />

observation alone explains the observed freeze tolerance is difficult<br />

to determine. Selected cryptobiotic species of tardigrades, nematodes<br />

as well as some freeze-tolerant insects tolerate as much as 80% of<br />

the body water being converted into ice (Westh and Kristensen,<br />

1992; Ramløv and Westh, 1993; Wharton and Block, 1997;<br />

Hengherr et al., 2009).<br />

The apparent morphological difference between the two stages<br />

is similarly relevant in regard to the observed freeze tolerance. The<br />

P1 stage is formed from an incomplete molt in which both the mouth<br />

and cloaca become sealed by cuticular thickenings (see Fig.1), and<br />

the gut content is often shed prior to this transition. In nature, ice<br />

nucleation can be initiated by a wide range of exogenous substances<br />

(Wharton and Worland, 1998). Consequently, the additional layer<br />

of cuticle could increase the capacity to avoid inoculative freezing<br />

in animals in the P1 stage, as has been demonstrated for eggs of<br />

the nematode Panagrolaimus davidi Timm 1971 (see Wharton,<br />

1994). Indeed, the extensive capacity for supercooling in both the<br />

active and P1 stage, along with the additional layer of cuticle and<br />

the clearing of gut contents in P1, would indicate that H. crispae<br />

preferentially seek to avoid internal ice formation. Nevertheless,<br />

animals in the P1 stage tolerate internal ice formation for both shorter<br />

and longer periods of time.<br />

Volume and osmoregulation<br />

Our results indicate that active-stage H. crispae is the most tolerant<br />

of changes in external salinity. Specimens from the population from<br />

Nipisat Bay, Greenland exhibit an increased tolerance towards<br />

concentrated SW solutions as compared with animals from Vellerup<br />

Vig, Denmark, suggesting that the ability to tolerate large increases<br />

in salinity is potentiated by living in a more-exposed habitat. The<br />

observed volume regulatory response of active-stage H. crispae<br />

during hypo- and hypertonic treatments differs in a significant way.<br />

When exposed to the hypo-osmotic solutions, the initial increase in<br />

total body volume was regulated to a new steady state, which was<br />

not significantly different from the control condition. In fact, a new<br />

steady-state value was demonstrated after merely 4h immersion in<br />

the external medium of 10ppt. Conversely, during acute exposure<br />

to concentrated seawater (40ppt), a partial recovery to normal levels<br />

was demonstrated; however, total body volume remained<br />

significantly different from the control condition even after 48h<br />

immersion. These data suggest that the body volume of H. crispae<br />

is more tightly regulated during exposure to dilute as compared with<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


2810<br />

K. A. <strong>Halberg</strong> and others<br />

more concentrated saltwater solutions. Interestingly, this observation<br />

seems reflected in an evolutionary context. According to our<br />

previous study, Halobiotus has evolved within the freshwater genus<br />

Isohypsibius, thus potentially explaining the enhanced volume<br />

regulatory response during exposure to dilute media (Møbjerg et<br />

al., 2007).<br />

When submitted to osmotic shock of 10 ppt and 40 ppt,<br />

respectively, our data show that H. crispae experience few<br />

limitations in terms of motility. Active-stage specimens from the<br />

Nipisat population even retain activity when exposed to a gradual<br />

salinity increase to 60ppt (Fig.3). However, upon direct transfer to<br />

an extreme seawater dilution of 2ppt, animal activity is markedly<br />

reduced, probably due to the pronounced increase in hydrostatic<br />

pressure and concomitant reduction in hemolymph osmotic pressure.<br />

Proper locomotory function in tardigrades relies on the hydrostatic<br />

pressure of the body cavity (Kinchin, 1994); thus, maintaining an<br />

appropriate body volume is essential to normal coordination of<br />

movement. In addition, as the membrane potentials of animal cells<br />

are highly dependent upon extracellular ionic strength (Spyropoulos<br />

and Teorell, 1968), the concomitant changes in hemolymph osmotic<br />

pressure could influence animal motility due to inhibition of neuromuscular<br />

activity. Indeed, the fact that a significant number of<br />

animals were observed passive during exposure to 2ppt, while<br />

animals remained largely unaffected during exposure to 10ppt, in<br />

spite of experiencing comparable average changes in body volume<br />

(compare Fig.4B and Fig.5B after 0.5h immersion), suggests that<br />

not only total body volume but also hemolymph osmolality is an<br />

important factor in maintaining locomotory functions.<br />

Exposing H. crispae to severe osmotic stress reveals that this<br />

species is a euryhaline osmoconformer, in which the hemolymph<br />

osmotic pressure is largely governed by the external environment.<br />

However, when analyzing hemolymph osmotic pressure at steady<br />

state following 48h exposure to the various salinity treatments as<br />

a function of the external osmolality, an interesting pattern emerges.<br />

H. crispae maintains a large osmotic pressure gradient between the<br />

internal and external environment, thus distinctly hyper-regulating<br />

during all investigated salinity treatments – albeit markedly less in<br />

concentrated seawater. This would imply a large water turnover in<br />

this animal, with osmotic water uptake being balanced by the<br />

excretion of dilute urine.<br />

Hyperosmoregulation is known in other euryhaline invertebrates.<br />

The crayfish Procambarus clarkii Girard 1852 (Arthropoda)<br />

produces highly dilute urine and is a strong hyperosmoregulator in<br />

freshwater (Sarver et al., 1994). However, the excreted urine<br />

becomes progressively more concentrated in media of higher ionic<br />

strength and is nearly isoosmotic when exposed to an external<br />

concentration of 750mOsmkg –1 , at which P. clarkii cease to hyperregulate<br />

(Sarver et al., 1994). Moreover, similar osmoregulatory<br />

responses have been reported from nematodes (Fusé et al., 1993;<br />

Forster, 1998). Indeed, the internal osmolality of the parasitic<br />

nematode Pseudoterranova decipiens (Krabbe, 1878) was<br />

maintained 90mOsmkg –1 above that of the external environment<br />

during exposure to media of widely varying osmolality (Fusé et al.,<br />

1993).<br />

Three glands positioned at the transition zone between the<br />

midgut and rectum of eutardigrades are traditionally ascribed an<br />

osmoregulatory function. The term used for these structures, i.e.<br />

Malpighian tubules, was introduced more than a century ago (Plate,<br />

1889). The positional conformity of the Malpighian tubules in<br />

eutardigrades and in hexapods has been used as a strong argument<br />

in favor of a homology between these structures (Greven, 1982;<br />

Møbjerg and Dahl, 1996). However, at present, no functional data<br />

exist relating the Malpighian tubules of tardigrades to an<br />

osmoregulatory role. Nevertheless, several detailed morphological<br />

investigations of the tubules support the hypothesis. These studies<br />

have provided ultrastructural data, which are in agreement with an<br />

active transporting epithelium involved in solute and fluid transport<br />

(Greven, 1979; Weglarska, 1987a; Weglarska, 1987b; Møbjerg and<br />

Dahl, 1996; Peltzer et al., 2007). As holds for insects, the Malpighian<br />

tubules of tardigrades are considered secretion–reabsorption kidneys.<br />

In light of the ultrastructural data available on tardigrade Malpighian<br />

tubules, it seems reasonable to assume that the first steps in urine<br />

formation take place across initial segment cells, characterized by<br />

a conspicuous basal labyrinth, numerous mitochondria and an<br />

enlarged apical surface. There is, moreover, ultrastructural support<br />

for assigning the tardigrade rectum an osmoregulatory function<br />

(Dewel and Dewel, 1979). A possible mode of urine formation was<br />

outlined by Dewel and Dewel (Dewel and Dewel, 1979). They<br />

suggest that isoosmotic urine produced by the Malpighian tubules<br />

is modified in the rectum through the active reabsorption of solutes,<br />

leading to the excretion of hypo-osmotic urine. Our data seem in<br />

favor of such a mechanism of urine formation. Interestingly,<br />

preliminary and unpublished data (H.R. research group) on<br />

Richtersius coronifer (Richters 1903) indicate that this species also<br />

remains hyperosmotic during exposures to a range of external<br />

salinities and it is therefore likely that hyper-regulation, and possibly<br />

hypo-osmotic urine formation, is a general feature of eutardigrades.<br />

However, until functional studies at the cellular and molecular level<br />

are performed, the exact mechanisms involved in osmoregulation<br />

in tardigrades remain to be elucidated.<br />

In conclusion, we show that the transition between the individual<br />

cyclomorphic stages of H. crispae is associated with profound<br />

changes in the physiology of the animal. Our results show that<br />

animals in the active stage tolerate large changes in the external<br />

osmotic pressure by regulating their total body volume and by<br />

enduring large concomitant changes in hemolymph osmotic<br />

pressure. H. crispae remains hyperosmotic at any investigated<br />

external salinity, suggesting that this species is a strong hyperregulator.<br />

Our study is the first to provide evidence for the volume<br />

and osmoregulatory capacity in Tardigrada. Whereas animals in the<br />

active stage are intolerant of freezing, the P1 stage is demonstrated<br />

to be freeze tolerant. The relatively low crystallization temperature<br />

reveals that extensive supercooling of the body fluids takes place<br />

during cooling and that no physiologically relevant ice-nucleating<br />

agents are present.<br />

We would like to thank the crew onboard the research vessel R/W Porsild (Arctic<br />

Station, Qeqertarsuaq, Greenland) who made the collection at the type locality,<br />

Nipisat, possible. Funding came from the Carlsberg Foundation and from the<br />

2008 Faculty of Science, University of Copenhagen Freja-Programme.<br />

DSC<br />

MDP<br />

RVD<br />

SW<br />

T c<br />

T m<br />

LIST OF ABBREVIATIONS<br />

differential scanning calorimetry<br />

melting point depression<br />

regulatory volume decrease<br />

seawater<br />

crystallization temperature<br />

melting temperature<br />

REFERENCES<br />

Block, W. (1991). To freeze or not to freeze Invertebrate survival of sub-zero<br />

temperatures. Funct. Ecol. 5, 284-290.<br />

Convey, P. and McInnes, S. J. (2005). Exceptional tardigrade-dominated ecosystems<br />

in Ellsworth Land, Antarctica. Ecology 86, 519-527.<br />

Dewel, R. A. and Dewel, W. C. (1979). Studies on the tardigrades. J. Morphol. 161,<br />

79-110.<br />

Forster, S. J. (1998). Osmotic stress tolerance and osmoregulation of intertidal and<br />

subtidal nematodes. J. Exp. Mar. Biol. Ecol. 224, 109-125.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


Cyclomorphosis in tardigrades<br />

2811<br />

Fusé, M., Davey, K. G. and Sommerville, R. I. (1993). Osmoregulation in the<br />

parasitic nematode Pseudoterranova decipiens. J. Exp. Biol. 175, 127-142.<br />

Garey, J. R., Krotec, M., Nelson, D. R. and Brooks, J. (1996). Molecular analysis<br />

supports a tardigrade-arthropod association. Invertebr. Biol. 115, 79-88.<br />

Giribet, G., Carranza, S., Baguña, J., Riutort, M. and Ribera, C. (1996). First<br />

molecular evidence of the existence of a Tardigrada+Arthropoda clade. Mol. Biol.<br />

Evol. 13, 76-84.<br />

Greven, H. (1979). Notes on the structure of vasa Malpihgii in the eutardigrade<br />

Isohypsibius augusti (Murray, 1907) Zesz. Nauk. Uniw. Jagiel. Prace Zool. 25, 87-<br />

95.<br />

Greven, H. (1982). Homologues or analogues A survey of some structural patterns in<br />

Tardigrada. In Proceedings of the Third International Symposium on the Tardigrada<br />

(ed. D. R. Nelson), pp. 55-76. Johnson City, TN: East Tennessee State University<br />

Press.<br />

Grøngaard, A., Pugh, P. J. A. and McInnes, S. (1999). Tardigrades, and other<br />

cryoconite biota, on the Greenland ice sheet. Zool. Anz. 238, 211-214.<br />

Guidetti, R., Boschini, D., Altiero, T., Bertolani, R. and Rebecchi, L. (2008).<br />

Diapause in tardigrades: a study of factors involved in encystment. J. Exp. Biol. 211,<br />

2296-2302.<br />

Hengherr, S., Heyer, A. G., Köhler, H. R. and Schill, R. O. (2008). Trehalose and<br />

anhydrobiosis in tardigrades: evidence for divergence in response to dehydration.<br />

FEBS J. 275, 281-288.<br />

Hengherr, S., Worland, M. R., Reuner, A., Brümmer, F. and Schill, R. O. (2009).<br />

Freeze tolerance, supercooling pionts and ice formation: comparative studies on the<br />

subzero temperature survival of limno-terrestrial tardigrades. J. Exp. Biol. 212, 802-<br />

807.<br />

Horikawa, D. D., Sakashita, T., Katagiri, C., Watanabe, M., Kikawada, T.,<br />

Nakahara, Y., Hamada, N., Wada, S., Funayama, T., Higishi, S. et al. (2006).<br />

Radiation tolerance in the tardigrade Milnesium tardigradum. Int. J. Radiat. Biol. 82,<br />

843-848.<br />

Jönson, K. I. and Schill, R. O. (2007). Induction of Hsp70 by dessication, ionizing<br />

radiation and heat-shock in the eutardigrade Richtersius coronifer. Comp. Biochem.<br />

Physiol. 146B, 456-460.<br />

Kinchin, I. M. (1994). The Biology of Tardigrades, pp. 1-186. London: Portland Press.<br />

Kristensen, R. M. (1982). The first record of cyclomorphosis in Tardigrada based on a<br />

new genus and species from Arctic meiobenthos. Z. Zool. Syst. Evol. Forsch. 20,<br />

249-270.<br />

Kristiansen, J. and Westh, P. (1991). Freezing behaviour of multilamellar vesicles in<br />

0.9% sodium chloride. Cryo Letters 12, 167-176.<br />

Lee, R. E. (1991). Principles of insect low temperature tolerance. In Insects at Low<br />

Temperatures (ed. R. E. Lee and D. L. Denlinger), pp. 17-46. New York: Chapman<br />

and Hall.<br />

Mallatt, J. M., Garey, J. R. and Shultz, J. W. (2004). Ecdysozoan phylogeny and<br />

Baysian inference: first use of nearly complete 28S and 18S rRNA gene sequences<br />

to classify the arthropods and their kin. Mol. Phylogenet. Evol. 31, 178-191.<br />

Møbjerg, N. and Dahl, C. (1996). Studies on the morphology and ultrastructure of the<br />

Malpighian tubules of Halobiotus crispae Kristensen, 1982 (Eutardigrada). Zool. J.<br />

Linn. Soc. 116, 85-99.<br />

Møbjerg, N., Jørgensen, A., Eibye-Jacobsen, J., <strong>Halberg</strong>, K. A., Persson, D. and<br />

Kristensen, R. M. (2007). New records on cyclomorphosis in the marine<br />

eutardigrade Halobiotus crispae (Eutardigrada: Hypsibiidae). J. Limnol. 66 Suppl. 1,<br />

132-140.<br />

Nicholajsen, H. and Hvidt, Å. (1994). Phase behaviour of the system trehalose-NaCl-<br />

Water. Cryobiology 31, 199-205.<br />

Pelzer, B., Dastych, H. and Greven, H. (2007). The osmoregulatory/excretory organs<br />

of the glacier-dwelling eutardigrade Hypsibius klebelsbergi Mihelcic, 1959<br />

(Tardigrada). Mitt. Hamb. Zool. Mus. Inst. 104, 61-72.<br />

Plate, L. H. (1889). Beiträge zur Naturgeschichte der Tardigraden. Zool. Jb. Anat.<br />

Ontog. 3, 487-550.<br />

Ramløv, H. and Westh, P. (1992). Survival of the cryptobiotic tardigrade Adorybiotus<br />

coronifer during cooling to –196°C: effect of cooling rate, trehalose level and short<br />

term pre-acclimation. Cryobiology 29, 125-130.<br />

Ramløv, H. and Westh, P. (1993). Ice formation in the freeze tolerant alpine weta<br />

Hemideina maori Huttton (Orthoptera; Stenopelmatidae). Cryo Letters 14, 169-176.<br />

Ramløv, H. and Westh, P. (2001). Cryptobiosis in the Eutardigrade Adorybiotus<br />

(Richtersius) coronifer: tolerance to alcohols, temperature and de novo protein<br />

synthesis. Zool. Anz. 240, 517-523.<br />

Ramløv, H., Wharton, D. A. and Wilson, P. W. (1996). Recrystalization in a freezingtolerant<br />

Antarctic nematode Panagrolaimus davidi, and an alpine weta Hemideina<br />

maori (Orthoptera; Stenopelmatidae). Cryobiology 33, 607-613.<br />

Sarver, R. G., Flynn, M. A. and Holliday, C. W. (1994). Renal Na, K-ATPase and<br />

osmoregulation in the crayfish Procambarus clarkii. Comp. Biochem. Physiol. 107A,<br />

349-356.<br />

Schill, R. O., Steinbrück, G. H. B. and Köhler, H. R. (2004). Stress gene (hsp70)<br />

sequences and quantitative expression in Milnesium tardigradum (Tardigrada) during<br />

active and cryptobiotic stages. J. Exp. Biol. 207, 1607-1613.<br />

Spyropoulos, C. S. and Teorell, T. (1968). The dependence of nerve membrane<br />

potentials upon extracellular ionic strength. Proc. Natl. Acad. Sci. USA 60, 118-125.<br />

Weglarska, B. (1987a). Studies on the excretory system of Isohypsibius granulifer<br />

Thulin (Eutardigrada). In Biology of Tardigrades: Selected Symposia and<br />

Monographs 1 (ed. R. Bertolani), pp. 15-24. Modena: U.Z.I. Mucchi.<br />

Weglarska, B. (1987b). Morphology and ultrastructure of the excretory system in<br />

Dactylobiotus dispar (Murray) (Eutardigrada). In Biology of Tardigrades: Selected<br />

Symposia and Monographs 1 (ed. R. Bertolani), pp. 25-33. Modena: U.Z.I. Mucchi.<br />

Westh, P. and Kristensen, R. M. (1992). Ice formation in the freeze-tolerant<br />

eutardigrades Adorybiotus coronifer and Amphibolus nebulosus studied by<br />

differential scanning calorimetry. Polar Biol. 12, 693-699.<br />

Wharton, D. A. (1994). Freezing avoidance in the eggs of the Antarctic nematode<br />

Panagrolaimus davidi. Fundam. Appl. Nematol. 17, 239-243.<br />

Wharton, D. A. and Block, W. (1997). Differential scanning calorimetry studies on an<br />

Antarctic nematode (Panagrolaimus davidi) which survives intracellular freezing.<br />

Cryobiology 34, 114-121.<br />

Wharton, D. A. and Worland, M. R. (1998). Ice nucleating activity in the freezingtolerant<br />

Antarctic nematode Panagrolaimus davidi. Cryobiology 36, 279-286.<br />

Zachariassen, K. E. (1985). Physiology of cold tolerance in insects. Physiol. Rev. 65,<br />

799-831.<br />

THE JOURNAL OF EXPERIMENTAL BIOLOGY


2012

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!