pmc logo imageJournal ListSearchpmc logo image
Logo of biochemjJournal URL: redirect3.cgi?&&auth=0lKr_22Xsf-K5K1wbOyvtUhDZwCrjEtCtSfoyLkuW&reftype=publisher&artid=1175106&article-id=1175106&iid=120253&issue-id=120253&jid=74&journal-id=74&FROM=Article|Banner&TO=Publisher|Other|N%2FA&rendering-type=normal&&http://www.biochemj.org/bj/currentissue.htm
Biochem J. 2005 July 15; 389(Pt 2): 297–306.
Published online 2005 July 5. Prepublished online 2005 April 5. doi: 10.1042/BJ20050305.
PMCID: PMC1175106
Structure, function and molecular adaptations of haemoglobins of the polar cartilaginous fish Bathyraja eatonii and Raja hyperborea
Cinzia Verde,* M. Cristina De Rosa, Daniela Giordano,* Donato Mosca, Donatella De Pascale,* Luca Raiola,* Ennio Cocca,* Vitale Carratore,* Bruno Giardina,1 and Guido Di Prisco*
*Institute of Protein Biochemistry, C.N.R., Via Marconi 12, I-80125 Naples, Italy
†Institute of Biochemistry and Clinical Biochemistry and C.N.R. Institute of Chemistry of Molecular Recognition, Catholic University, I-00168 Rome, Italy
1To whom correspondence should be addressed (email bgiardina/at/rm.unicatt.it).
The nucleotide sequence data reported will appear in DDBJ, EMBL, GenBank® and GSDB Nucleotide Sequence Databases under the accession numbers AY772716 (Bathyraja eatonii α-chain nucleotide sequence), AY772717 (Bathyraja eatonii β-chain nucleotide sequence), AY773131 (Raja hyperborea α-chain nucleotide sequence) and AY773132 (Raja hyperborea β-chain nucleotide sequence). The amino acid sequence data reported will appear in Swiss-Prot Protein Database under the accession numbers P84216 (Bathyraja eatonii α-chain amino acid sequence) and P84217 (Bathyraja eatonii β-chain amino acid sequence).
Received February 17, 2005; Revised March 24, 2005; Accepted March 31, 2005.
Abstract
Cartilaginous fish are very ancient organisms. In the Antarctic sea, the modern chondrichthyan genera are poorly represented, with only three species of sharks and eight species of skates; the paucity of chondrichthyans is probably an ecological consequence of unusual trophic or habitat conditions in the Southern Ocean. In the Arctic, there are 26 species belonging to the class Chondrichthyes. Fish in the two polar regions have been subjected to different regional histories that have influenced the development of diversity: Antarctic marine organisms are highly stenothermal, in response to stable water temperatures, whereas the Arctic communities are exposed to seasonal temperature variations. The structure and function of the oxygen-transport haem protein from the Antarctic skate Bathyraja eatonii and from the Arctic skate Raja hyperborea (both of the subclass Elasmobranchii, order Rajiformes, family Rajidae) is reported in the present paper. These species have a single major haemoglobin (Hb 1; over 80% of the total). The Bohr-proton and the organophosphate-binding sites are absent. Thus the haemoglobins of northern and southern polar skates appear functionally similar, whereas differences were observed with several temperate elasmobranchs. Such evidence suggests that, in temperate and polar habitats, physiological adaptations have evolved along distinct pathways, whereas, in this case, the effect of the differences characterizing the two polar environments is negligible.
Keywords: Antarctic, Arctic, Bohr effect, haemoglobin, phosphate binding, skate
Abbreviations: 2,3-BPG, 2,3-biphosphoglycerate; MALDI–TOF, matrix-assisted laser-desorption ionization–time-of-flight; PDB, Protein Data Bank
INTRODUCTION

Oxygen carriers are ancient proteins, probably evolved from enzymes that protected the organism against oxygen toxicity [1]. When multicellular organisms increased in size and complexity, their surface/volume ratios diminished, and simple diffusion of oxygen across the body wall was inadequate to reach all cells. The evolution of simple oxygen-binding proteins into multisubunit circulating proteins, in combination with the advent of circulatory systems, made the transport of oxygen from the periphery of the organism to cells possible. The evolution of the tetramer, including duplication and differentiation of the globin gene into α and β genes, formation of subunit interfaces, and quaternary-structure allosteric changes took place within a relatively short period between the branching points of hagfish and lampreys from cartilaginous fish. The separation of sharks from bony fish occurred near the α/β gene duplication [2].

In Antarctica, the chondrichthyan genera are poorly represented in the modern fauna, with only three species of sharks and eight species of skates [3], representing approx. 4% of the total fauna. The late-Cretaceous and late-Eocene fossils from Seymour and James Ross Islands are proof that many modern chondrichthyan genera inhabited the temperate waters of the Weddellian Province [4].

On the other hand, Notothenioidei are a large group of marine teleosts largely endemic to the Antarctic Ocean. Indirect indications suggest that notothenioids appeared in the early Tertiary, filling the ecological void on the shelf left by most of the other fish fauna (which became locally extinct during maximal glaciation), and began to diversify in the middle Tertiary. Reduced competition and increasing isolation favoured speciation. Notothenioids fill a varied range of ecological niches normally occupied by taxonomically diverse fish communities in temperate waters.

In this environment, the low metabolic demand and the high oxygen concentration reduce the need for Hb. The vast majority of species of the dominant suborder Notothenioidei have a single Hb, sometimes accompanied by a minor component [5]. Of the 213 species living on the shelf or upper slope of the Antarctic continent, 96 are notothenioids [6].

Why are there so few chondrichthyans in Antarctica? There is no obvious physiological reason for their scarcity. The freezing point of the blood plasma of Squalus acanthias (spiny dogfish) is −1.95 °C [7]. Since sea water freezes at −1.86 °C, this would seem to provide adequate protection against freezing, provided there is no contact with ice. Perhaps the scarcity of chondrichthyans in the modern fauna is an ecological consequence of unusual trophic or habitat conditions in the Southern Ocean [8]. In addition, the reduced diversity of teleost fishes in the Antarctic midwaters may have restricted the entry of sharks into the ecosystem.

Besides the shark species, in the Antarctic region, there are two species of Raja and six species of Bathyraja, the dominant family south of 60°S [9]. Bathyraja lacks a fossil record. The genus includes 45 extant species distributed worldwide and is supposed to have originated at least 100 million years ago [10]. Stehmann [10] proposed the following vicariance hypothesis: “During the Jurassic (213–144 million years ago), the Bathyraja lineage had an extensive latitudinal distribution in the shelf waters of most continental areas. Waters at this time were warm and homogenous throughout the world. Tectonic processes split the Bathyraja lineage into at least two stocks, one in the north-eastern Pacific and one off the South America/Antarctic component of Gondwana. At the same time waters were cooling toward the poles, and latitudinal climatic zones were developing”.

In comparing the taxonomic composition of the Arctic and Antarctic faunas, the most obvious difference is that no single group in the Arctic dominates the fauna as do the Antarctic notothenioids, where the species endemism is 88% for the benthic fauna and rises to 97% when only notothenioids are considered [11]. Species endemism for marine fish in the Arctic is 20–25% [12]. In the Arctic region, there are 26 species belonging to the class Chondrichthyes, approx. 6.3% of the total fauna.

Hbs of cartilaginous fish have been studied less extensively than those of teleosts. Only temperate elasmobranch Hbs have been investigated so far. One difficulty that presumably hindered the elucidation of structure–function relationships is the high genetic polymorphism. When the X-ray structures of two cartilaginous fish Hbs were solved [13,14], the quaternary structure and its change upon ligand binding appeared to be preserved. The proximal effect linking the quaternary-structure change to the haems was also preserved, but the distal effect, especially the role of valine E11, was altered. The Bohr-proton and organophosphate-binding sites were absent, indicating that, in such cases, the stereochemical mechanisms other than the proximal effect have evolved independently in the different species.

The present paper reports the first molecular characterization of the oxygen-transport haemoproteins of two skates living in polar habitats, namely Antarctic Bathyraja eatonii (Eaton's skate) and Arctic Raja hyperborea. The blood of the two species was found to contain two components. To assess the possible differential contribution of Hb within the evolutionary response to the two polar environments, the primary structure, the oxygenequilibrium properties and the thermodynamic features of the major component Hb 1 of both species were investigated. The primary structures of B. eatonii and R. hyperborea Hb 1 reveal significant substitutions of many residues that are generally conserved in many Hbs and in human HbA. These are likely to be related to the functional properties of these Hbs, i.e. absence of pH and organophosphate regulation, indicating that teleost and cartilaginous fish Hbs have independently evolved stereochemical mechanisms to regulate ligand binding to the haems. Moreover, similar to all cartilaginous fish, Hb 1 of B. eatonii and R. hyperborea (i) lacks helix D in the β-subunits, and (ii) has relatively low co-operativity (h≈1.6–1.8) [15]. Thermodynamic analysis indicates that the oxygenation enthalpy change (ΔH) maintains a rather constant absolute value in the pH range 6.6–8.7 in both species.

Molecular modelling was used to characterize the haem environment and to explain the lack of Bohr and organophosphate effects in comparison with the solved X-ray structures of the two cartilaginous fish [13,14].

EXPERIMENTAL

Materials
Toyopearl Super Q-650S was from TosoHaas, Mono P HR 5/20 was from Amersham Biosciences, trypsin (EC 3.4.21.4) treated with L-1-tosylamide-2-phenylethylchloromethylketone was from Cooper Biomedical, endoproteinase Asp-N and Glu-C (sequencing grade) were from Roche, 4-vinylpyridine was from Sigma, dithiothreitol was from Fluka, sequanal-grade reagents were from Applied Biosystems, HPLC-grade acetonitrile was from Lab-Scan Analytical, oligonucleotides were from MWG, and Taq DNA polymerase was from EuroClone. All other reagents were of the highest purity commercially available.

Hb purification
Specimens of the two skates were collected by bottom trawling from the research vessel L.M. Gould near Low and Braband Islands in the Palmer Archipelago (B. eatonii) and from the research vessel Jan Mayen near the coast of Greenland (R. hyperborea).

Blood samples were drawn by cardiac puncture by means of heparinized syringes. Haemolysates were prepared as described in [16]. Purification of Hb 1 was achieved by FPLC anion-exchange chromatography, with Toyopearl Super Q-650S (B. eatonii), and Amersham Biosciences Mono P HR 5/20 columns (R. hyperborea). The Hb-containing pooled fractions were dialysed against 10 mM Hepes, pH 7.7. All steps were carried out at 0–5 °C. Hb solutions were stored in small aliquots at −80 °C until use. For oxygen binding, aliquots of a solution of CO Hb 1 were stored at −80 °C before use within a maximum of 7 days. For each experiment, one aliquot was thawed, converted into the oxy form by exposure to light and oxygen, and was used immediately. For this purpose, the Hb solution was placed in an ice bath, and the gas phase was made 100% in oxygen. With gentle agitation, the Hb was illuminated with a light source (Sylvania Model SG-50 with a DWY lamp). No oxidation was detectable spectrophotometrically in the Hb solution, indicating that final Met-Hb formation was negligible (<2%).

Amino acid sequencing
Alkylation of thiol groups with 4-vinylpyridine, and tryptic, Asp-N and Glu-C digestions were carried out as described in [1719]. Globins and peptides were purified by reverse-phase HPLC on a micro-Bondapak-C18 column (0.39 cm×30 cm; Waters) as described in [17]. Cleavage of Asp-Pro bonds was performed on polybrene-coated glass-fibre filters in 70% (v/v) methanoic (formic) acid, for 24 h at 42 °C [20]. Asp-Pro-cleaved α-globins were treated with o-phthalaldehyde before sequencing [21] in order to block the non-proline N-terminus and reduce the background. Sequencing was performed using an Applied Biosystems Procise 492 automatic sequencer, equipped with on-line detection of phenylthiohydantoin amino acids. Multiple alignment of globins was performed with CLUSTALW [22].

Cloning and sequence analysis of globin cDNAs
Total RNA was isolated from spleen using TRI® Reagent (Sigma–Aldrich), as described in [23]. First-strand cDNA synthesis was performed according to the manufacturer's instructions (Promega) using an oligo(dT)-adaptor primer in both species. The α- and β-globin cDNAs were amplified by PCR using oligonucleotides designed on the N-terminal regions as direct primers and the adaptor primer as the reverse primer. Amplifications were performed with 2.5 units of Taq DNA polymerase, 5 pmol each of the above primers and 0.20 mM dNTPs buffered with 670 mM Tris/HCl, pH 8.8, 160 mM ammonium sulphate, 0.1% Tween 20 and 1.5 mM MgCl2. The PCR program consisted of 30 cycles of 1 min at 94 °C, 1 min at temperatures between 42 and 54 °C, and 1 min at 72 °C, and ending with a single cycle of 10 min at 72 °C. The cloning of the N-terminal regions was obtained by 5′ RACE (rapid amplification of cDNA ends) using the Marathon™ cDNA Amplification Kit (BD Biosciences) and two internal primers. Amplified cDNA was purified and ligated in the pDrive vector (Qiagen). Escherichia coli cells (strain DH5α) were transformed with the ligation mixtures. Standard molecular-biology techniques [24] were used in the isolation, restriction and sequence analysis of plasmid DNA. Both strands of the cloned cDNA fragments underwent automated sequencing.

MS
The molecular mass of S-pyridylethylated α- and β-chains and of peptides (less than 10 kDa) was measured by MALDI–TOF (matrix-assisted laser-desorption ionization–time-of-flight) MS on a PerSeptive Biosystems Voyager-DE Biospectrometry Workstation. Analyses were performed on pre-mixed solutions prepared by diluting samples (final concentration, 5 nmol·ml−1) in 4 vol. of matrix, namely 10 mg·ml−1 sinapinic acid in 30% acetonitrile containing 0.3% trifluoroacetic acid (globins), and 10 mg·ml−1 α-cyano-4-hydroxycinnamic acid in 60% acetonitrile containing 0.3% trifluoroacetic acid (peptides).

Oxygen binding
Haemolysate stripping was carried out as described in [25]. Oxygen equilibria were measured in 100 mM Hepes in the pH range 6.2–8.7, at 2 °C and 10 °C (keeping the pH variation as a function of temperature in due account) at a final Hb concentration of 0.5–1.0 mM on a haem basis. An average S.D. of ±3% for values of p50 (partial pressure of oxygen required to saturate 50% of the haems) was calculated; experiments were performed in duplicate. In order to obtain stepwise oxygen saturation, a modified gas diffusion chamber was used, coupled to cascaded Wösthoff pumps for mixing pure nitrogen with air [26]. pH was measured with a Radiometer BMS Mk2 thermostatically controlled electrode. Sensitivity to chloride was assessed by adding NaCl to a final concentration of 100 mM. The effects of ATP, GTP and inositol hexakisphosphate were measured at a final ligand concentration of 3 mM, namely a large excess over tetrameric Hb concentration. Oxygen affinity (p50) and co-operativity (h) were calculated from the linearized Hill plot of log S/(1−S) against log pO2 at half saturation; S denotes fractional oxygen saturation.

The overall ΔH (kcal·mol−1; 1 kcal=4.184 kJ), corrected for the heat of oxygen solubilization (−3 kcal·mol−1), was calculated by the integrated van't Hoff equation:

equation M1

Molecular modelling
A specific BLAST sequence search of the Protein Data Bank (PDB) using default parameters was used to identify candidate parent structures of B. eatonii and R. hyperborea Hb 1 for homo-logy modelling. Next, sequence alignments were performed using CLUSTALW for sequences of the temperate skate Dasyatis akajei (red stingray) Hb (PDB code 1CG5) [13], the shark Mustelus griseus (spotless smooth-hound) Hb (PDB code 1GCV) [14], human HbA (PDB code 2HHB) [27] and the Antarctic teleost Trematomus bernacchii (emerald rockcod) Hb (PDB code 1HBH) [28].

Models of α- and β-chains of B. eatonii and R. hyperborea Hb 1 were built using both the multiple alignments obtained by CLUSTALW and the single structure of D. akajei Hb, which, among the chosen sequences, shows the highest identity with the target (62% and 59% for α-chain, 56% and 58% for β-chain of B. eatonii and R. hyperborea Hb 1 respectively). Three different models for both alignments were made using the comparative modelling program MODELER [29] as implemented in InsightII (Accelrys). The zone of β-chain around β50, which corresponds to the region of removal of helix D in cartilaginous fish, was optimized further (loop modelling in MODELER). The structural validity of the models of α- and β-chains of B. eatonii and R. hyperborea Hb 1 was assessed by several criteria. A representative model from each set was selected by reference to the MODELER objective function (F, molecular probability density function violation), which describes the degree of fit of the model to the input structural data used in its construction, as well as by use of the structure verification program PROCHECK [30]. The co-ordinates of the generated models were then subjected to the Verify3D algorithm [31] using the Verify3D Structure Evaluation Server (available at http://www.doe-mbi.ucla.edu/Services/Verify_3D.html) to identify regions of improper folding. Sequence–structure compatibility was also assessed with the ProsaII program [32].

After addition of hydrogen atoms, the quaternary structures of B. eatonii and R. hyperborea Hb 1 were created using InsightII (Accelrys) and CHARMM force field for energy minimization. Energetically favourable positions for the binding of organophosphates in B. eatonii Hb 1, R. hyperborea Hb 1 and D. akajei Hb were evaluated using the docking program GROUP implemented in GRID version 21 [33]. The program GRID determines the interaction energy of a ‘probe’, representing a chemical fragment, at each point of an orthogonal grid which encloses the structure of a target macromolecule. A grid which surrounds each protein structure exceeding it by 5 Å (1 Å=0.1 nm) in each dimension and a grid spacing of 1.0 Å were selected. The standard GRID energy function and parameters were used to calculate the interaction energy. The program GROUP, by fitting the maps generated by GRID for the probes which all together represent the structure of ATP, positions the atoms of the ligand at favourable places on the protein.

RESULTS AND DISCUSSION

Purification of Hbs and separation of globins
Ion-exchange chromatography of the haemolysate (results not shown) showed two components in both skates. Purification of Hb 1 was achieved by gradient elution with Tris/HCl, pH 7.6 (B. eatonii) and Tris/HCl, pH 7.6, containing NaCl (R. hyperborea). The first peak corresponded to Hb 1, the second to Hb 2 (contaminated by Hb 1). The approximate amount ratio was 80:20. The literature suggests that the various components present in cartilaginous-fish haemolysates are often structurally and functionally similar. It appears that such haemolysates contain products of genetic polymorphism, which can be regarded as a single type of Hb which merely exhibits microheterogeneity in the amino acid sequence.

The separation of the globins of B. eatonii and R. hyperborea Hb 1 was obtained by reverse-phase HPLC (results not shown). HPLC, SDS/PAGE (15% polyacrylamide) and MS indicated the presence of four different subunits in the haemolysate of both species. However, five cDNAs for α-chains and three cDNAs for β-chains were identified in B. eatonii (E. Cocca and G. di Prisco, unpublished work); in keeping with the hypothesis of genetic polymorphism, these sequences differ from one another at very few positions.

Primary structure
The amino acid sequences of the α- and β-chains of Hb 1 of B. eatonii and R. hyperborea are shown in Figure 1. The primary structure of Hb 1 of B. eatonii was established by alignment of tryptic, Asp-N and Glu-C peptides (results not shown), and on the basis of nucleotide sequences (E. Cocca and G. di Prisco, unpublished work) to confirm the amino acid sequence and to complete the C-terminal region of the β-chains. The primary structure of Hb 1 of R. hyperborea was established by nucleotide sequencing, using primers designed on sequence stretches. The N-terminus of the α-chains in both species was available for Edman degradation. The sequence-deduced molecular masses of globins were 15680 Da for α- and 15668 Da for β-chains (B. eatonii), and 15564 Da for α- and 15399 Da for β-chains (R. hyperborea). These values are in agreement with MALDI–TOF MS.
Figure 1Figure 1
Amino acid sequences of the α- and β-chains of B. eatonii and R. hyperborea Hb 1

B. eatonii Hb 1 has 141 amino acid residues in the α- and β-chains; R. hyperborea has 139 amino acid residues in the α-chain and 141 in the β-chain. By comparison with the α-chain of human HbA, one deletion (αGH3) and one insertion (between αCD8 and αCD9) were found in B. eatonii Hb 1, and three deletions (αB1, αB2 and αGH3) and one insertion (between αCD8 and αCD9) were found in R. hyperborea Hb 1. The α-chains of both fish, similar to human HbA and of the temperate skate D. akajei Hb, have free valine at the N-terminus. In both β-chains, four residues are missing between helices C and E, corresponding to helix D (Figure 1).

Remarkably, in B. eatonii Hb 1, there is a serine residue at position H21 (143β), compared with histidine in mammals and lysine or arginine in other fish (except trout Hb 1, which also has serine and lacks organophosphate regulation); R. hyperborea Hb 1 has alanine at H21 (143β) (Figure 1).

The presence of histidine at NA2 (2β) in R. hyperborea is not exceptional for elasmobranchs. It also occurs in Hbs of the sharks S. acanthias [34] and Heterodontus portusjacksoni (Port Jackson shark) [35,36]. The episodic occurrence of histidine in NA2 (2β) in elasmobranchs suggests that it may be a phylogenetically primitive character.

The alignment of the polar α- and β-chains with those of HbA and of Hbs of D. akajei, the shark M. griseus and the Antarctic bony fish T. bernacchii is shown in Figures 2 and 3. The sequences of the two chains of Hbs from polar cartilaginous Hbs have comparatively higher identity with other temperate cartilaginous Hbs than with polar teleosts (Table 1). The sequence identity between Hbs of the Arctic and Antarctic skates is even higher (70–73%).

Figure 2Figure 2
Sequence features of α- (A) and β- (B) chains of B. eatonii Hb 1
Figure 3Figure 3
Sequence features of α- (A) and β- (B) chains of R. hyperborea Hb 1
Table 1Table 1
Sequence identities of the two chains of Hb of various fish species compared with B. eatonii and R. hyperborea

The amino acid sequences of the α- and β-chains of B. eatonii Hb 2 were also established (results not shown). The identity with the homologous chains of Hb 1 was over 95%, supporting the microheterogeneity hypothesis in cartilaginous fish Hbs.

Oxygen-binding properties
B. eatonii and R. hyperborea Hb 1 are devoid of Bohr effect at 10 and 2 °C, both in the absence and the presence of effectors (Figures 4A, 4C, 5A and 5C respectively). The chloride-independent part of the Bohr effect of human HbA largely arises from the β-subunit C-terminal residue His146 β(HC3), which forms salt bridges with Asp94 β(FG1) through the imidazole group and with Lys40 α(C5) across the α1β2 contact through the carbonyl group [37]. In B. eatonii and R. hyperborea Hb 1, positions β(HC3) and α(C5) are occupied by the same residues, but Asp94 β(FG1) is replaced by leucine in B. eatonii Hb 1 and by threonine in R. hyperborea Hb 1, uncommon in other fish Hbs (Figure 1). Moreover, both species lack organophosphate regulation (see below).
Figure 4Figure 4
Oxygen-equilibrium isotherms (Bohr effect) (A, C) and subunit co-operativity (B, D) as a function of pH of B. eatonii Hb 1 at 10 °C (A, B) and at 2 °C (C, D)
Figure 5Figure 5
Oxygen-equilibrium isotherms (Bohr effect) (A, C) and subunit co-operativity (B, D) as a function of pH of R. hyperborea Hb 1 at 10 °C (A, B) and at 2 °C (C, D)

The changes in oxygen-transport protein synthesis, whether up- or down-regulation or expression of a particular combination of gene products, are often considered to be a long-term response to developmental or environmental change. On the other hand, the regulation by allosteric effectors (protons and organophosphates) is thought to be responsible for more immediate short-term perturbations.

At first glance, the p50 values of polar skate Hb 1 appeared to be low when compared with some other cartilaginous Hbs, in both the absence and the presence of allosteric effectors (Table 2). However, this difference should be considered in the light of the fact that Hbs of some cartilaginous fish display the Bohr effect. Therefore, at alkaline pH, their affinity is much higher, and, at low pH, it becomes much lower [13,14]. Moreover, if the affinities are compared, taking the respective environmental temperatures and physiological pH into account, this difference becomes insignificant. As shown in Figures 4(B), 4(D), 5(B) and 5(D), co-operativity was relatively low in both species, reaching a maximum of approx. 1.8, as in all cartilaginous fish [15], indicating that a highly co-operative oxygen carrier is not needed.

Table 2Table 2
Oxygen affinity values (p50) for polar and temperate cartilaginous fish

Enthalpy change
Oxygen-binding equilibria were investigated in the range 2–10 °C in both Hbs (Figure 6). In Hb 1 of both fish, where the Bohr effect and phosphate regulation is absent, the ΔH, which includes the heat of oxygen solubilization and the heats of processes linked to oxygen binding, such as proton and anion dissociation, kept a rather constant absolute value as a function of pH, in line with the absence of the endothermic contributions due to the Bohr effect and ATP regulation. The evolution of polar fish appears to have often favoured a decrease in the temperature sensitivity of Hb oxygen affinity, surprisingly similar to that observed in some species whose lifestyle is highly different, and which must adapt to large temperature fluctuations, e.g. tuna [25]. Thus these Hbs do not require significant amounts of energy during the oxygenation–deoxygenation cycle. This energy-saving mechanism may facilitate Hb function in the constantly low temperature of sea water.
Figure 6Figure 6
Overall oxygenation enthalpy change of B. eatonii (A) and R. hyperborea (B) Hb 1

Temperature-dependence, which is governed by the associated overall enthalpy change, is an important feature of the reaction of Hbs with oxygen. Heat absorption and release can be considered to be physiologically relevant modulating factors, similar to hetero- and homo-tropic ligands. The affinity and co-operativity of all Hb systems are strictly correlated with the extent of heat absorption accompanying oxygenation. In human HbA, the apparent overall ΔH is more exothermic at alkaline pH values, where the Bohr effect is not operative and the endothermic contribution of the Bohr protons is abolished. For fish, relying upon Hbs with constant ΔH values may thus be a frequent evolutionary strategy of molecular adaptation to extreme life conditions.

Molecular modelling
In order to gain additional insight into the functional properties of B. eatonii and R. hyperborea Hb 1, structural models were built by homology modelling, relying on crystal co-ordinates of closely related Hbs. A three-dimensional model of the structure of both Hbs was built on the basis of homology with four X-ray structures of Hb from the PDB, namely the 1.6 Å resolution structures of Hb of the temperate skate D. akajei, the 2.0 Å resolution structure of Hb of the shark M. griseus, the 2.2 Å resolution structure of Hb of the Antarctic teleost T. bernacchii and the 1.74 Å resolution structure of human HbA.

A second model was built based on the single structure of D. akajei Hb; these proteins share approx. 60% sequence identity (Table 1), and, in the case of B. eatonii Hb 1, there are no insertions or deletions in the alignment. Sequence alignments are shown in Figures 2 and 3. From the three models based on the multiple alignment and the three models based on D. akajei Hb, a single representative model was chosen for each, in accordance with the MODELER objective function and PROCHECK stereochemical assessment. The stereochemistry of the selected models was acceptable in both cases. The models show no residues within disallowed regions, and only one or two residues in the generously allowed regions of the Ramachandran plot, indicating correct stereochemistry.

The overall quality of the structure was also determined by parameters that assess residue environments and atomic contacts. As shown in Figures 7(A) and 7(B), the three/one-dimensional scores of the models generated by Verify3D for α- and β-chains respectively are always positive and are similar to those obtained with the template structure of temperate D. akajei (PDB code 1CG5).

Figure 7Figure 7
Verify3D plots

The reliability of the selected models was also tested by the ProsaII program, which calculates the Cβ-Cβ pair interaction energy for each residue in the sequence, producing smooth energy plots with negative values for correctly folded proteins. The ProsaII trace for candidate models of α- and β-chains of B. eatonii Hb 1 had no positive regions, indicating no misfolding, with a so-called Z-score or normalized energy values of −6.89 and −6.72 (calculated for α-chains from multiple and single alignment respectively), and −6.64 and −6.74 (calculated for β-chains from multiple and single alignment respectively), comparing favourably with those of D. akajei Hb (−5.67 and −6.04 for α- and β-chains respectively). Analogously, the Z-score values for R. hyperborea were comparable to those of Hb crystals of the temperate skate (−6.63 and −6.53 for α-chains from multiple and single alignment respectively, and −6.23 and −6.21 for β-chains from multiple and single alignment respectively).

The haem environment
Analysis of the predicted models of B. eatonii and R. hyperborea Hb 1 showed no significant mutations in the haem region of both α- and β-chains compared with D. akajei Hb and human HbA. It is worth noting that removal of D helix in β-subunits has been reported not to affect significantly the conformation of the distal pocket [38].

In the proximal side of both α- and β-haems, the replacement of leucine F7 of HbA with lysine in D. akajei, B. eatonii and R. hyperborea Hb 1 (Figure 1) is worth mentioning.

In the α-haem distal cavity, glycine E3 of human HbA and D. akajei Hb is replaced by proline in B. eatonii Hb 1, and by histidine in R. hyperborea Hb 1, thus modifying the hydrogen-binding interaction pattern of distal histidine. In the β-chains, it is worth noting that two polar residues (histidine FG4 and lysine E3), involved in hydrogen-bonding with proximal and distal histidine of HbA respectively, are replaced by glycine in D. akajei Hb, B. eatonii and R. hyperborea Hb 1 (Figure 1).

Organophosphate-binding site
The lack of effector modulation displayed by Hb 1 of polar skates can be explained by a molecular modelling study predicting the favourable binding sites for ATP. In both polar species, lacking organophosphate regulation, the ATP-binding site was absent. A model for the binding site of ATP in bony fish Hbs has been proposed [39]. In this model, the same residues [lysine β(EF6), histidine β(H21), histidine β(NA2) and β N-terminal α-amino group] which bind 2,3-BPG (2,3-biphosphoglycerate) in HbA take part in the binding of ATP. In bony fish Hb, glutamate or aspartate replaces histidine at NA2 and arginine replaces histidine at H21, but the former residue can be modelled to accept a hydrogen bond from adenine and the latter to participate in the binding of phosphate. With respect to bony fish Hb, in D. akajei, B. eatonii and R. hyperborea Hbs, lysine β(EF6) is replaced by an acidic residue and arginine β(H21) is replaced by serine in B. eatonii Hb 1 and by alanine in R. hyperborea Hb 1; in βNA2, there is lysine residue in B. eatonii and a histidine residue in R. hyperborea. Thus skate Hbs cannot bind ATP with the same interaction pattern characteristic of bony fish Hbs. However, in D. akajei Hb, Chong et al. [13] suggested, as a candidate binding site for ATP, the region of arginine β(H13) and lysine β(G6) in the central cavity just inside the 2,3-BPG-binding site of human HbA.

By using the program GRID, we succeeded in demonstrating that ATP is indeed bound to D. akajei Hb as hypothesized [13], whereas the ATP site is absent in the polar skate Hb 1, since the replacement of Arg130 β(H13) with leucine in both Hbs does prevent organophosphate binding. Figure 8 shows the predicted binding site for ATP in D. akajei Leu130 β(H13) of B. eatonii and R. hyperborea Hbs, which disallows the binding within the protein cavity, is superimposed.

Figure 8Figure 8
Organophosphate-binding site in D. akajei Hb

Bohr effect
The replacement of Asp94 β(FG1) of HbA with leucine in B. eatonii Hb and threonine in R. hyperborea Hb (Figure 1) causes the loss of salt bridges which are essential for the Bohr effect in human HbA. D. akajei Hb compensates the effects of the substitution of glutamate for Asp94 β(FG1), with the hydrogen bond between His141 β(HC3) and Asn139 β(HC1) likely to be responsible for part of the Bohr effect displayed by Hb of the temperate skate [13]. In contrast, in B. eatonii and R. hyperborea Hb, the presence of glycine in HC1 gives rise to an arrangement which does not allow hydrogen bonding formation.

Concluding remarks
Globins of higher vertebrates have eight helices, designated A–H from the N-terminus, but all known vertebrate α-globins lack helix D. The absence of helix D in α-subunits is considered to be the most distinct conserved feature which distinguishes modern α- and β-globins. The widespread distribution of helix-D-containing globins throughout plants, bacteria and ancestral vertebrates suggests that helix D was present in the very early ancestral protein, but was lost in the α-globin of vertebrates.

A feature which distinguishes cartilaginous from teleost Hbs is the lack of helix D in the β subunits of the former. In β-globins, it was shown previously that the presence or absence of helix D does not affect assembly into co-operative tetramers [40]. Since recombinant human Hbs engineered with β-subunits without helix D and α-subunits having helix D show a small change in oxygen affinity, the lack of helix D in the β-subunit is thought not to exert a large functional effect. B. eatonii and R. hyperborea Hb 1 do not have helix D in either α- or β-subunits, as well as the sharks S. acanthias [34], H. portusjacksoni [35,36] and M. griseus [14], and the temperate skate D. akajei [13].

The very high sequence similarity observed in B. eatonii and R. hyperborea Hbs may reflect a common origin of polar skates, but may also suggest that the primary structure in polar fish may be related to the development of cold adaptation.

Oxygen delivery to the tissues of B. eatonii and R. hyperborea occurs in the virtual absence of pH or phosphate-linked reductions in Hb 1 affinity.

These features resemble those of ancestral oxygen carriers, and also of Hbs of several amphibians, of trout Hb I and of abnormal Hiroshima Hb (histidine βHC3 replaced by aspartate), but are not unique to polar elasmobranchs. In fact, previous studies on temperate elasmobranch Hbs have shown that their functional properties do not always include well-developed co-operativity or proton and organophosphate regulation. For instance, in Torpedo marmorata Hb, the replacement of aspartate βFG4 with lysine [41] causes the loss of the salt bridge crucial for the Bohr effect in human HbA [39]; this mutation may be responsible for the lack of Bohr effect, as the Hb of Torpedo nobiliana binds oxygen with low co-operativity (h<1), and shows no appreciable pH-dependence or response to organophosphates [42]. Elasmobranchs may have diverged from the mammalian line before stabilization of the homotropic and heterotropic interactions typical of mammalian Hbs.

In contrast, some cartilaginous fish (e.g. D. akajei and M. griseus) display a modest Bohr effect and significant ATP regulation. Such regulation originated in the common ancestor of the subclass Elasmobranchii and all other jawed vertebrates; ATP was the first organophosphate regulator of Hb function [43]. The switch from ATP to 2,3-BPG regulation may have been a consequence of curtailed oxidative phosphorylation in erythrocytes of higher vertebrates such as mammals.

In summary, Hbs of polar skates appear functionally different from those of several temperate cartilaginous fish. On the other hand, the close functional similarity (also reflected in the very high sequence identity) in the two species, each living in one of the polar environments, is of interest. The Hb systems of teleosts thriving at the two poles are far from showing the sequence identities found in skates. Antarctic notothenioids, with very few exceptions, are sedentary bottom dwellers, and have a single major Hb usually displaying strong Bohr and Root effects. Among Arctic teleosts, characterized by higher biodiversity, the number of pelagic and migratory species is instead very abundant, the Hb multiplicity is higher, and the various components appear to be functionally distinct [44,45]. Although both are cold, the Arctic and Antarctic habitats differ in many aspects, e.g., in the Arctic, the range of temperature variations is wider and isolation is less pronounced. These differences appear to be minimized in the case of two benthic sedentary skates, unlikely to disperse across wide latitude and temperature gradients and living in microhabitats with similar temperatures; this similar lifestyle may have channelled the Hb functional mechanisms (Bohr effect and ATP regulation) towards coinciding evolutionary pathways. Conversely, it is conceivable that the wide environmental differences between temperate and polar waters have given rise to far more extensive differences in physiological adaptations to the respective environments than those caused by the two polar environments, which do differ from each other, but to a lesser extent.

Physiological studies provide tantalizing insights into the possible mechanisms used by polar fish in response to low temperature. The comparison of structure and function of proteins from cold-adapted and non-cold-adapted species is an additional powerful tool. In fact, it will allow us to gain insights into the extent to which strategies of cold-adaptation are similar or vary in different phylogenies, and to what extent extreme environments require specific adaptations or simply select for more generalist or phenotypically plastic lifestyles.

Acknowledgments

This study is in the framework of the Italian National Programme for Antarctic Research (PNRA), the Arctic Strategic Programme of the Italian National Research Council, the SCAR (Scientific Committee on Antarctic Research) programmes Ecology of the Antarctic Sea Ice Zone (EASIZ), Evolutionary Biology of Antarctic Organisms (EVOLANTA) and Evolution and Biodiversity in the Antarctic (EBA), and the 2003 cruise TUNU-I (Greenland).

References
1.
Terwilliger, N. B. Functional adaptations of oxygen-transport proteins. J. Exp. Biol. 1998;201:1085–1098. [PubMed]
2.
Dickerson, R. E.; Geis, I. Menlo Park: The Benjamin/Cummings Publishing Company, Inc.; 1983. Haemoglobin: Structure, Function, Evolution, and Pathology.
3.
Gon, O.; Heemstra, P. C. Grahamstown: JLB Smith Institute of Ichthyology; 1990. Fishes of the Southern Ocean.
4.
Eastman, J. T. San Diego: Academic Press; 1993. Antarctic Fish Biology: Evolution in A Unique Environment.
5.
di Prisco, G. Molecular adaptation of Antarctic fish haemoglobins. Fishes of Antarctica: a Biological Overview. In: di Prisco G., Pisano E., Clarke A. , editors. Milan: Springer; 1998. pp. 339–353.
6.
Eastman, J. T. Antarctic notothenioid fishes as subjects for research in evolutionary biology. Antarct. Sci. 2000;12:276–287.
7.
Watts, R.; Watts, D. C. Nitrogen metabolism in fishes. In: Florkin M., Scheer B. T. , editors. Chemical Zoology. New York: Academic Press; 1974. pp. 369–446.
8.
Grande, L.; Eastman, J. T. A review of Antarctic ichthyofaunas. Palaeontology. 1986;29:113–137.
9.
Stehmann, M.; Burkel, D. L. Rajidae. Fishes of the Southern Ocean. In: Gon O., Heemstra P. C. , editors. Grahamstown: JLB. Smith Institute of Ichthyology; 1990. pp. 86–97.
10.
Stehmann, M. Notes on the systematics of the rajid genus Bathyraja and its distribution in the world oceans. In: Uyeno T., Arai R., Taniuchi T., Matsuura K. , editors. Indo-Pacific Fish Biology: Proceedings of the Second International Conference on Indo-Pacific Fishes. Tokyo: Ichthyological Society of Japan; 1986. pp. 261–268.
11.
Andriashev, A. P. A general review of the Antarctic bottom fish fauna. In: Kullander, S. O.; Fernhalm, B., editors. Proceedings of the Fifth Congress of European Ichthyology. Stockholm: Swedish Museum of Natural History; 1987. pp. 357–372.
12.
Briggs, J. C. New York: McGraw–Hill; 1974. Marine Zoogeography.
13.
Chong, K. T.; Miyazaki, G.; Morimoto, H.; Oda, Y.; Park, S. Y. Structures of the deoxy and CO forms of haemoglobin from Dasyatis akajei, a cartilaginous fish. Acta Crystallogr. Sect. D Biol. Crystallogr. 1999;55:1291–1300. [PubMed]
14.
Naoi, Y.; Chong, K.-T.; Yoshimatsu, K.; Miyazaki, G.; Tame, J.-R.-H.; Park, S.-Y.; Adachi, S.; Morimoto, H. The functional similarity and structural diversity of human and cartilaginous fish haemoglobins. J. Mol. Biol. 2001;307:259–270. [PubMed]
15.
Brittain, T. The Root effect. Comp. Biochem. Physiol. 1987;86B:473–481.
16.
D'Avino, R.; di Prisco, G. Antarctic fish haemoglobin: an outline of the molecular structure and oxygen binding properties. I. Molecular structure. Comp. Biochem. Physiol. 1988;90B:579–584.
17.
D'Avino, R.; di Prisco, G. Haemoglobin from the Antarctic fish Notothenia coriiceps neglecta. 1. Purification and characterisation. Eur. J. Biochem. 1989;179:699–705. [PubMed]
18.
Tamburrini, M.; Brancaccio, A.; Ippoliti, R.; di Prisco, G. The amino acid sequence and oxygen-binding properties of the single haemoglobin of the cold-adapted Antarctic teleost Gymnodraco acuticeps. Arch. Biochem. Biophys. 1992;292:295–302. [PubMed]
19.
Tamburrini, M.; D'Avino, R.; Fago, A.; Carratore, V.; Kunzmann, A.; di Prisco, G. The unique haemoglobin system of Pleuragramma antarcticum, an Antarctic migratory teleost: structure and function of the three components. J. Biol. Chem. 1996;271:23780–23785. [PubMed]
20.
Landon, M. Cleavage at aspartyl-prolyl bonds. Methods Enzymol. 1977;47:145–149. [PubMed]
21.
Brauer, A. W.; Oman, C. L.; Margolies, M. N. Use of o-phthalaldehyde to reduce background during automated Edman degradation. Anal. Biochem. 1984;137:134–142. [PubMed]
22.
Thompson, J. D.; Higgins, D. G.; Gibson, T. J. CLUSTALW: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994;22:4673–4680. [PubMed]
23.
Chomczynski, P.; Sacchi, N. Single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction. Anal. Biochem. 1987;162:156–159. [PubMed]
24.
Sambrook, J.; Fritsch, E. F.; Maniatis, T. 2nd edn. Cold Spring Harbor: Cold Spring Harbor Laboratory Press; 1989. Molecular Cloning: a Laboratory Manual.
25.
Tamburrini, M.; Condò, S. G.; di Prisco, G.; Giardina, B. Adaptation to extreme environments: structure–function relationships in Emperor penguin haemoglobin. J. Mol. Biol. 1994;237:615–621. [PubMed]
26.
Weber, R. E.; Jensen, F. B.; Cox, R. P. Analysis of teleost haemoglobin by Adair and Monod–Wyman–Changeux models: effects of nucleoside triphosphates and pH on oxygenation of tench haemoglobin. J. Comp. Physiol. 1987;B157:145–152.
27.
Fermi, G.; Perutz, M. F.; Shaanan, B.; Fourme, R. The crystal structure of human deoxyhaemoglobin at 1.74 Å resolution. J. Mol. Biol. 1984;175:159–174. [PubMed]
28.
Ito, N.; Komiyama, N. H.; Fermi, G. Structure of deoxyhaemoglobin of the Antarctic fish Pagothenia bernacchii with an analysis of the structural basis of the Root effect by comparison of the liganded and unliganded haemoglobin structures. J. Mol. Biol. 1995;250:648–658. [PubMed]
29.
Sali, A.; Blundell, T. L. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 1993;234:779–815. [PubMed]
30.
Laskowski, R. A.; MacArthur, M. W.; Moss, D. S.; Thornton, J. M. PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Crystallogr. 1993;26:283–291.
31.
Luthy, R.; Bowie, J. U.; Eisenberg, D. Assessment of protein models with three-dimensional profiles. Nature (London). 1992;356:83–85. [PubMed]
32.
Sippl, M. Recognition of errors in three-dimensional structures of proteins. Proteins. 1993;17:355–362. [PubMed]
33.
Goodford, P. J. A computational procedure for determining energetically favorable binding sites on biologically important macromolecules. J. Med. Chem. 1985;28:849–857. [PubMed]
34.
Aschauer, H.; Weber, R. E.; Braunitzer, G. The primary structure of the haemoglobin of the dogfish shark (Squalus acanthias): antagonistic effects of ATP and urea on oxygen affinity of an elasmobranch haemoglobin. Biol. Chem. Hoppe Seyler. 1985;366:589–599. [PubMed]
35.
Nash, A. R.; Fisher, W. K.; Thompson, E. O. Haemoglobins of the shark, Heterodontus portusjacksoni. II. Amino acid sequence of the α-chain. Aust. J. Biol. Sci. 1976;29:73–97. [PubMed]
36.
Fisher, W. K.; Nash, A. R.; Thompson, E. O. Haemoglobins of the shark, Heterodontus portusjacksoni. III. Amino acid sequence of the β-chain. Aust. J. Biol. Sci. 1977;30:487–506. [PubMed]
37.
Perutz, M. F. The stereochemical mechanism of the cooperative effects in haemoglobin revisited. Annu. Rev. Biophys. Biomol. Struct. 1998;27:1–34. [PubMed]
38.
Whitaker, T. L.; Berry, M. B.; Ho, E. L.; Hargrove, M. S.; Phillips, G. N. J.; Komiyama, N. H.; Nagai, K.; Olson, J. S. The D-helix in myoglobin and in the β subunit of haemoglobin is required for the retention of haem. Biochemistry. 1995;34:8221–8226. [PubMed]
39.
Perutz, M. F.; Brunori, M. Stereochemistry of cooperative effects in fish and amphibian haemoglobins. Nature (London). 1982;299:421–426. [PubMed]
40.
Komiyama, N. H.; Shih, D. T.; Looker, D.; Tame, J.; Nagai, K. Was the loss of the D helix in α globin a functionally neutral mutation? Nature (London). 1991;352:349–351. [PubMed]
41.
Huber, F.; Braunitzer, G. The primary structure of electric ray haemoglobin (Torpedo marmorata). Bohr effect and phosphate interaction. Biol. Chem. Hoppe Seyler. 1989;370:831–838. [PubMed]
42.
Bonaventura, J.; Bonaventura, C.; Sullivan, B. Haemoglobin of the electric Atlantic torpedo, Torpedo nobiliana: a cooperative haemoglobin without Root effect. Biochim. Biophys. Acta. 1974;371:147–154. [PubMed]
43.
Coates, M. L. Haemoglobin function in the vertebrates: an evolutionary model. J. Mol. Evol. 1975;6:285–307. [PubMed]
44.
Verde, C.; Carratore, V.; Riccio, A.; Tamburrini, M.; Parisi, E.; di Prisco, G. The functionally distinct haemoglobins of the Arctic spotted wolffish Anarhichas minor. J. Biol. Chem. 2002;277:36312–36320. [PubMed]
45.
Verde, C.; Parisi, E.; di Prisco, G. Comparison of Arctic and Antarctic teleost haemoglobins: primary structure, function and phylogeny. Antarct. Sci. 2004;16:59–69.