From the Laboratorium voor Fysiologie, Katholieke
Universiteit Leuven, Campus Gasthuisberg, Herestraat 49, B-3000 Leuven,
Belgium and the
Department of Biochemistry and Molecular
Biology, School of Medicine, University of Louisville, Louisville,
Kentucky 40292
Received for publication, August 23, 2000, and in revised form, October 20, 2000
![]() |
ABSTRACT |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
We have compared the primary sequence
and enzymatic properties of the sarcoplasmic reticulum
Ca2+-ATPases from a cold-tolerant frog Rana
sylvatica with those of a closely related cold-intolerant frog,
Rana clamitans. Sarcoplasmic reticulum isolated from
leg muscles of both species contains a major protein (~100 kDa) that
reacts with a monoclonal antibody against sarco(endo)plasmic reticulum
Ca2+-ATPase type 1 (SERCA1). The apparent molecular
mass of R. sylvatica SERCA1 is 115 kDa, whereas that
of R. clamitans is 105 kDa. However, the deduced amino acid
sequences obtained from cDNAs do not indicate a difference in
molecular weight, thus suggesting post-translational protein
modification of R. sylvatica SERCA1. Comparison of the temperature dependence of both ATP hydrolysis and Ca2+
transport indicates that R. sylvatica SERCA1 exhibits
significantly lower activation energy below 20 °C and an ~2-fold
greater Ca2+-ATPase activity near 0 °C. Furthermore,
R. sylvatica SERCA1 exhibits simple Michaelis-Menten
kinetics with ATP and Ca2+ as opposed to the two-site ATP
kinetics and positive cooperativity with Ca2+ observed for
R. clamitans and mammalian SERCA1s. Cooperativity has been
linked to protein-protein interaction in SERCA1, and this property may
be altered in R. sylvatica SERCA1. Primary sequence comparison shows that R. sylvatica SERCA1 exhibits seven
unique amino acid substitutions, three of which are in the ATP binding domain. We also report for the first time the presence of alternative splicing in the frog, resulting in isoforms SERCA1a and SERCA1b. Thus,
it appears that the low temperature muscle contractility of R. sylvatica can be explained partially by significant functional and structural differences in SERCA1.
There has been considerable interest in the mechanism by which the
wood frog, Rana sylvatica, survives freezing. This frog is
one of the few vertebrate species that can be frozen solid and revived
upon thawing (1). R. sylvatica survives freezing by
producing very high cryoprotectant concentrations of glucose in vital
organs (2, 3). Furthermore, R. sylvatica mates and lays eggs
in water temperatures approaching 0 °C. Because R. sylvatica muscle functions efficiently near 0 °C, the
Ca2+ pump must also be functional at low temperatures.
However, although the effects of freezing/thawing on several enzymes
and physiological processes have been studied (4-6), no studies on the
wood frog skeletal muscle Ca2+-ATPase
(SERCA1)1 have been
reported. The kinetic properties and amino acid sequence of
SERCA1 from Rana esculenta (European green frog) have been reported (7, 8) and are quite similar to those of the rabbit. However,
R. sylvatica muscle could not function at near freezing temperatures if its Ca2+-ATPase had the same properties as
the rabbit because rabbit SERCA1 is almost totally inactive at 0 °C
(9). Kossler and Kuchler (10) concluded that the main cause of severe
inhibition of muscle contraction in the rat at 6 °C is the
temperature-dependent inhibition of Ca2+ uptake
and release. A similar conclusion was reached for another frog
Rana temporaria at 10 °C (11). Thus, SERCA1 is likely to have altered properties in the skeletal muscle of the cold-tolerant frog R. sylvatica.
There is a vast amount of knowledge concerning structure/function
relationships for SERCA1 because it is one of the most highly studied
membrane-associated proteins. Structure/function relationships have
recently been reviewed by MacLennan et al. (12). The protein was first purified in 1970 by MacLennan (13), and the sequence was
deduced from the cDNA by MacLennan et al. (14). The
complete three-dimensional structure has recently been obtained at
2.6-Å resolution by x-ray crystallographic analysis of protein
crystals (15). This recent advance has confirmed the three-dimensional folding of the domains that had previously been proposed based on
protein chemistry and analysis of two-dimensional arrays in membranes.
An astonishing proportion of the amino acid residues predicted to be
involved in SERCA1 function has been modified by site-directed
mutagenesis and the effects on the function of the expressed protein
were determined (12). However, there is no direct structural
information regarding temperature dependence of the pump.
In the present study, we have compared structural and functional
properties of SERCA1s from R. sylvatica with those of
Rana clamitans, a closely related frog living in the same
environment that is cold-sensitive and cannot function at 0 °C. Our
results indicate that there are significant differences in the kinetic properties of SERCA1 between these two species, including temperature dependence of ATP hydrolysis and Ca2+ transport, kinetics
with ATP, and cooperativity of Ca2+ kinetics. We also
report the complete amino acid sequences of these two SERCA1s.
Materials and Chemicals--
R. sylvatica and
R. clamitans were collected from local ponds in accordance
with the guidelines provided by the Institutional Animal Care and Use
Committee. The chemicals used for SDS-polyacrylamide gel
electrophoresis and immunoblotting were obtained from Bio-Rad. Chemiluminescence reagents were purchased from PerkinElmer Life Sciences. Monoclonal anti-SERCA1, clone VE121G9 was from Affinity BioReagents, Inc. (Golden, CO) as were the monoclonal anti-SERCA2 (MA3-910) clone IID8 and polyclonal anti-peptide anti-SERCA3 (PA1-910) antisera. All three antibodies exhibit cross-reactivity with several mammalian species. Restriction endonucleases were supplied by New
England Biolabs, Inc. (Beverly, MA). The Expand® Long
Distance PCR kit, which uses a mixture of Taq and
Pwo polymerases, was from Roche Molecular Biochemicals.
ALF® Express Sequencing kit and cyanin dATP-Labeling Mix
as well as cyanin-labeled sequencing primers were purchased from
Amersham Pharmacia Biotech. Plasmid preparations were carried out using Qiagen columns (Qiagen, Hilden, Germany). Restriction fragments used
for subcloning procedures were gel-purified with the QIAEX II Gel
Extraction kit (Qiagen). pGEM®-T Easy vector (Promega,
Madison, WI) was chosen for direct subcloning of all 5' and 3' RACE-PCR
products. Gene-specific oligonucleotides were from Life Technologies,
Inc. Other molecular biology grade chemicals were purchased from Sigma.
Preparation and Analysis of Sarcoplasmic Reticulum (SR)--
Leg
muscles were removed from euthanized R. sylvatica captured
in late February and from R. clamitans captured in
late March, both soon after emergence from hibernation. SR membranes
were purified from the homogenized muscle by sucrose density
centrifugation using the procedure described by Bull et al.
(16). SDS-polyacrylamide gel electrophoresis was carried out according
to Laemmli (17), and immunoblotting was done as described earlier (18).
Ca2+-ATPase activity was assayed using a system with ATP
hydrolysis coupled to NADH oxidation by pyruvate kinase and lactate
dehydrogenase (19). In addition to the coupling components, the assay
contained 10 mM Tes buffer at pH 7.4, 0.1 M
KCl, and 5.0 mM MgCl2. ATP was prepared as a
1:1 MgATP complex. ATP and Ca2+ dependence of
Ca2+-ATPase activity were assayed at 30 °C in a Bio-Tek
Powerwave X 96-well plate reader, whereas the temperature dependence of Ca2+-ATPase activity was determined in a Cary 219 spectrophotometer. Ca2+-ATPase activity was calculated from
the difference in the rate of ATP hydrolysis in the presence of various
concentrations of Ca2+ and 2 mM EGTA. For
Ca2+ dependence, free [Ca2+] was calculated
taking into account the concentrations of total Ca2+, EGTA,
Mg2+, ATP, temperature and pH using the Winmaxc
program available on the world wide web (Chris Patton, Stanford
University). Ca2+ transport was measured using the
Ca2+-sensing dye, Rhod2, as described previously (20). The
buffer components were the same as the components listed above for the Ca2+-ATPase assay except that MgCl2 was
eliminated, and 2.5 µM Rhod2 was added. ATP (5 mM) was added as the MgATP complex. The assay was carried
out at various temperatures in a PerkinElmer LS50B Luminescence
Spectrometer with excitation at 553 nm and emission at 576 nm. Before
the addition of membranes (a final concentration of 0.1 mg/ml), 0.5 nmol aliquots of CaCl2 were added to calibrate the
assay and allow for the determination of Ca2+ transport
rates. The addition of BrA23187 at the end of each assay released all
sequestered Ca2+ (20). Kinetic data were analyzed using
Enzfitter© software (version 1.05), whereas curve fitting was
accomplished using the Slide Write Plus© software (version 5.0).
Protein concentration was estimated using bicinchoninic acid (BCA
Protein Assay kit, Pierce) with bovine serum albumin as a standard.
Construction of Wood and Green Frog cDNA
Libraries--
Uncloned adaptor-ligated double-stranded (ds) cDNA
libraries were constructed using the MarathonTM cDNA amplification
kit from CLONTECH (Palo Alto, CA) according to
manufacturer's instructions. Total RNA was isolated from either an
entire wood frog (weighing 1 g) at an early age or a green frog
skeletal muscle (weighing 1 g) using the guanidinium
thiocyanate/CsCl purification method (21). 10 µg of total RNA was
used for each first-strand synthesis. The MarathonTM cDNA synthesis
primer (52-mer) was extended with Moloney murine leukemia virus reverse
transcriptase for 1 h at 42 °C, after which
second-strand synthesis was carried out by adding an enzyme mixture
containing RNase H, Escherichia coli DNA polymerase I, and
E. coli DNA ligase for 2 h at 16 °C. A blunt ending
of the ds-cDNAs was performed with T4 DNA polymerase at 16 °C
for 45 min. ds-cDNAs were purified by
phenol/CHCl3/isoamyl alcohol (25:24:1) extraction
precipitated by 95% ethanol in the presence of 4 M
ammonium acetate. The pellet was washed with 80% ethanol and dissolved
in sterile water. The MarathonTM cDNA adaptor was ligated to
ds-cDNAs overnight at 16 °C using T4 DNA ligase.
Rapid Amplification of 3' and 5' cDNA Ends by PCR (3' and 5'
RACE-PCR)--
3' and 5' RACE-PCR amplifications were performed using
the adaptor-ligated ds-cDNAs as template DNAs, a mixture of
Taq and Pwo (with a proofreading activity)
polymerases, a common Marathon adaptor-primer 1, and an internal
gene-specific primer. For the 3' RACE-PCR, a gene-specific primer PH+
(5'-TCTGATAAGACTGGTACTCTGACCACC-3') was derived from the
nucleotide sequence coding for the phosphorylation site
(350SDKTGTLTT) of the R. esculenta
SERCA1a (7). The PH+ primer corresponds to nucleotides 1275-1301 in
the R. esculenta cDNA sequence (numbered according to
the sequence deposited under GenBankTM accession number
X63009) and to nucleotides 1251-1277 and 1254-1280, respectively, in
R. sylvatica and R. clamitans nucleotide sequences described in this study. However, note that the underlined base "T" of the R. esculenta is replaced by a "C" in
both R. sylvatica and R. clamitans cDNA
sequences. The 3' RACE-PCR amplifications with the PH+/adaptor primer 1 primer set were carried out for 30 cycles, each cycle consisting of 1 min of denaturation at 94 °C, 1 min of annealing at 60 °C, and 4 min of extension at 68 °C. PCR products were size-fractionated by
1% agarose gel electrophoresis, visualized by ethidium bromide
staining, gel-purified, and subcloned into the pGEM®-T
Easy vector. For each frog, several clones were partially sequenced. To
perform the 5' RACE-PCR reactions, an antisense primer (5'R Reverse Transcriptase-PCR Analysis of R. sylvatica SERCA1 Splice
Variants--
The primers used to analyze the possibility of an
alternative splicing event of the SERCA1 primary transcript are as
follows: a 5' primer, SPLICE+ (5'-GTAGAACAGTGGTTCGTTGTACTG-3')
corresponding to nucleotides 3093-3116 and a 3' primer, SPLICE Analysis of R. sylvatica and R. clamitans
SR--
SDS-polyacrylamide gel analysis of a Coomassie Blue-stained
gel is shown in Fig. 1A. The
major protein in the SR of both R. sylvatica and R. clamitans is ~100 kDa. Both preparations contain lesser amounts
of several other proteins. Immunoblot analysis with monoclonal
anti-SERCA1 reveals that the major 100-kDa proteins are indeed SERCA1
(Fig. 1B). Antibodies against SERCA2 and SERCA3 (see
"Experimental Procedures" for details) did not recognize any frog
proteins. SERCA1 in R. sylvatica exhibits a molecular mass
of 115 kDa, whereas that of R. clamitans is ~105 kDa.
Thus, the SERCA1s differ in apparent mass between the two frog species. This is not a general property because the other SR proteins evident in
Fig. 1A exhibit the same molecular weights. This difference in apparent mass could indicate the presence of post-translational modifications that affect the enzymatic properties of the proteins.
Kinetic analyses were carried out to determine whether there are
functional differences between the two SERCAs. The specific activities
at 37 °C of two preparations of R. sylvatica sarcoplasmic reticulum were 0.8 µmol (used in Figs. 1 and
2A) and 3.0 µmol of ATP
hydrolyzed/min/mg (used in Figs. 2B, 3 and 4), whereas the
specific activity of a single R. clamitans preparation was 2.7. These values are similar to those observed for rabbits (13).
Because we hypothesize that R. sylvatica SERCA1 is more
active at 0 °C than that of R. clamitans, we analyzed the
temperature dependence of SERCA1 Ca2+-ATPase activity. ATP
hydrolysis activity was determined at saturating ATP (10 mM) and Ca2+ (100 µM), and the
results were analyzed using the Arrhenius plot after calculating the
relative activities normalized to the maximum rate at 37 °C. As
shown in Fig. 2A, R. sylvatica SERCA1 has a higher relative activity at all temperatures tested, and activities for
the two frogs at the two lowest temperatures (6 and 2 °C
corresponding to 3.584 and 3.636 × 10
Fig. 2B shows a similar analysis for Ca2+
transport data. Between 13 and 30 °C, the activation energy for
Ca2+ transport is again higher for the green frog,
21.7 ± 0.6 versus 19.0 ± 0.1 kcal/mol (mean ± S.E.), yielding an extrapolated value for Ca2+ transport
in the wood frog at 0 °C that is 1.8-fold greater than in the green
frog. This result suggests a similar coupling of ATP hydrolysis to
Ca2+ transport over a wide range of temperatures.
Next, the dependence of ATP hydrolytic rate on ATP concentration was
determined. Fitted plots of wood and green frog kinetics are shown in
Fig. 3. The R. clamitans
SERCA1 data fit a two-site model with Km of 0.005 and 0.34 mM. The correlation coefficient (r2) for the two-site model is 0.999, whereas
attempting to fit the data to Michaelis-Menten kinetics yields a
relatively poor correlation coefficient of 0.77. Thus, the green frog
exhibits ATP kinetics quite similar to that described for rabbit SERCA1
(19). This behavior has been ascribed to half of the site reactivity in
a SERCA1 dimer or additional ATP binding sites on a single SERCA1 monomer (22). Interestingly, the R. sylvatica SERCA1 data
can be fitted with a simple single-site Michaelis-Menten model yielding a Km of 0.12 ± 0.03 mM (mean ± S.E.) (r2 = 0.94). Dean and Tanford (19)
demonstrated that the multiple ATP site behavior was lost upon monomer
formation of rabbit SERCA1 in C12E8, thus
suggesting the possibility that protein-protein interactions are
altered in R. sylvatica SR. The combination of higher
activity at a low temperature and lower apparent Km could result in an enhanced ability of R. sylvatica SERCA1
to function at low temperature, especially in the physiological range of ATP concentrations (0.5-5 mM).
The kinetics of stimulation of ATP hydrolysis by Ca2+ was
also analyzed, as shown in Fig. 4. As
indicated by the error bars in the plot, there is a significant
difference in the Ca2+ dependence of
Ca2+-ATPase activity for the two frogs at both low and high
Ca2+ concentrations. The similarity in activities near the
center of the plot indicate similar apparent affinities for
Ca2+ (0.5 µM). Fitting the plots to a model
of positive cooperativity yields a Hill coefficient of 1.5 ± 0.2 (mean ± S.D.) for the green frog indicating a positive
cooperativity, a model that has been well established for rabbit SERCA1
(23). In contrast, the wood frog data yield a Hill coefficient of
0.9 ± 0.2 (mean ± S.D.) indicating a lack of cooperativity.
This result suggests a loss of protein-protein interactions in R. sylvatica SERCA1 because the positive cooperativity of
Ca2+ binding to rabbit SERCA1 has been shown to be lost
when the protein is dissociated to monomers in
C12E8 (23).
Characterization of cDNA Clones Coding for Wood and Green
Frog SERCAs--
To obtain the nucleotide sequence of the two frog
SERCA cDNAs, we have employed the RACE-PCR strategy (24), which
allows the amplification of both 3' and 5' cDNA ends. To conduct
the 3' RACE reactions, a gene-specific oligonucleotide (PH+) was
designed from the phosphorylation site of the known sequence of the
previously cloned cDNA of the European frog (R. esculenta). For both frog SERCA cDNAs, 3' RACE-PCR products
(2.3 kilobase pairs) were obtained and sequenced. From the
obtained sequence, a new primer (5'R Amino Acid Sequence Analysis and Comparison with Other SERCA
Pumps--
The primary translation products deduced from the two
composite nucleotide sequences are 994 amino acids in length. The
predicted molecular masses of R. sylvatica and R. clamitans SERCA proteins are, respectively, 109.7 and 109.3 kDa.
In Fig. 6, the deduced amino acid
sequence of R. sylvatica SERCA is compared with the sequences of other SERCA pumps expressed in skeletal muscle from R. clamitans, R. esculenta, and rabbit. In
pairwise comparisons, R. sylvatica SERCA exhibits 98% (982 of 994 amino acids) overall identity with green frog SERCA and
95% (948 of 994 amino acids) and 88% (879 of 994 amino acids)
identity with R. esculenta and rabbit SERCA1a, respectively.
The deduced R. sylvatica SERCA sequence clearly matches that
of rabbit fast-twitch skeletal muscle isoform (SERCA1a) at two
characteristic points of deviation between rabbit SERCA1a and SERCA2a
(expressed in cardiac and slow-twitch muscles) isoforms. One of them is
represented by the COOH terminus (ARNYLEG994), whereas the
other consists of a 5-amino acid stretch (AAVGN510) located
COOH-terminal to Arg505. Therefore, based on these
considerations and the overall sequence identity, we suggest that the
nomenclature used for R. sylvatica and R. clamitans Ca2+-ATPases that are encoded by the cloned
cDNAs be SERCA1a as it was proposed earlier for R. esculenta (7). The hydrophobic putative transmembrane domains
(M1-M10), the phosphorylation site (residues 349-355), as well as
sites known to bind fluorescein isothiocyanate or the ATP
analogues FSBA and CIRATP in SERCA1 (27-29) are conserved in
R. sylvatica SERCA1a. There are five nonconservative amino
acid substitutions (Asn, Gln, Ala, Ser, and Leu) characteristic for
R. sylvatica (Fig. 6 and Table
I), respectively, at positions 26 (NH2 terminus region), 546, 547, 643 (nucleotide binding
domain), and 838 (transmembrane domain M7). In the other SERCA pumps,
these positions are occupied by highly conserved residues:
Pro26 or Leu26, Leu546,
Ser547, Gly643 and Met838.
Alternative Processing of the Wood Frog SERCA1 Primary
Transcript--
So far, no information suggesting the alternative
splicing of SERCA1 primary transcript in frog species has been
reported. We are now providing the first evidence for the expression of two SERCA1 splice variants in a frog species, R. sylvatica,
by means of the PCR strategy using SPLICE+ and SPLICE The results presented herein may provide at least a partial
explanation for the unusual low temperature muscle contractility of the
wood frog R. sylvatica. R. sylvatica SERCA1 at saturating levels of ATP and Ca2+ exhibits a 2-fold higher
Ca2+-ATPase activity and an estimated 1.5-fold higher
Ca2+ transport rate at 0 °C than R. clamitans
(Fig. 2, A and B). Furthermore, R. sylvatica SERCA1 at 0.5 mM ATP has an activity
1.6-fold higher than that of R. clamitans (Fig. 3).
Together, these differences in enzymatic activity may enable R. sylvatica SR to maintain muscle contractility near 0 °C in
contrast to R. clamitans. The fact that
excitation/contraction coupling is maintained in the wood frog at
0 °C signifies that very low Ca2+-ATPase activity
at this temperature (1.5% of the activity at 37 °C) is sufficient
for SR function. The published physiological experiments indicate that
in both rat and frog, loss of muscle function at low temperature
results from the diminution of Ca2+ release and re-uptake
(10, 11). However, it is possible that the lower activity of green frog
SERCA1 at 0 °C (0.7% of the activity at 37 °C) is not the
rate-limiting step that prevents muscle function at this temperature in
this frog.
Measurements of temperature dependence of rabbit SERCA1 activity in
native SR membranes indicate an apparent discontinuity at ~20 °C
(32, 33), and the data shown in Fig. 2A can also be fit with
two linear segments and a break at 20 °C. It was originally thought
that this discontinuity in temperature dependence of ATP hydrolysis was
a direct indicator of an abrupt change in the fluidity of the SR
membrane (33). Subsequent work indicated that a lipid phase transition
does not occur at 20 °C, and it was proposed that the discontinuity
was a result of changes in lipid cluster formation (34) or in a tightly
associated lipid annulus (35). In contrast, Dean and Tanford
(19) showed that the discontinuity at 20 °C was maintained
in the absence of phospholipid, which argues against any specific
effect of lipid on SERCA1 at 20 °C. Nevertheless, it is
clear that the activation energy for Ca2+-stimulated ATP
hydrolysis is higher for R. clamitans SERCA1 than for
R. sylvatica below 20 °C (28 versus 23.7 kcal/mol). These values are in the published range for rabbit, lobster,
scallop, and winter flounder (32, 36-38). The Arrhenius plots of
ATPase activity versus temperature for the cold-tolerant
scallop, lobster, flounder, and trout are similar to that of the
rabbit, but they show modest shifts to lower activation energies below
20 °C as is reported here for R. sylvatica.
Interestingly, trout SERCA1 showed a nearly constant activation energy
from 4 to 37 °C for phosphoenzyme formation from
phosphate resulting in much higher levels of this activity at 4 °C
than that observed for rabbit. The temperature dependence of
phosphoenzyme formation in lobster, scallop, winter flounder, or frog
has not been reported. Similarly, the cooperativity of kinetics with
Ca2+ and ATP observed to be different in R. sylvatica compared with R. clamitans or rabbit has not
been described for lobster, scallop, or winter flounder. Kalabokis and
Hardwicke (36) concluded that the lipid environment contributed to the
differences in properties observed between rabbit and scallop SERCA1
because scallop SERCA1 solubilized in C12E8
exhibited an Arrhenius plot similar to that of solubilized rabbit
SERCA1. However, in the trout it was implied that differences in the
temperature dependence of phosphoenzyme formation reside in the protein
itself and not in the lipid environment due to the inability of
C12E8 to alter the temperature dependence of
phosphoenzyme formation (37). Thus, the role of the phospholipid environment in altering temperature dependence of SERCA1 in the SR of
cold-tolerant species is not established. This is an area for future
work on R. sylvatica SR.
Several lines of evidence indicate that SERCA1 is aggregated in its
native SR and that this interaction has a functional significance. Intrinsic fluorescence, fluorescence energy transfer, chemical cross-linking, and analytical ultracentrifugation all indicate a
propensity for aggregation of SERCA1 (30). The positive cooperativity of Ca2+ binding has been proposed to be a result of
protein-protein interactions (23, 30). Similarly, Dean et
al. (19, 31) have shown that solubilization of SERCA1 with the
nonionic detergent C12E8 results in the
formation of an active monomer but that multiple site ATP kinetics is
lost in the monomer. Because cooperativity of Ca2+
activation and multiple site ATP kinetics are lost in R. sylvatica, it is possible that there is a change in
protein-protein interactions in comparison with R. clamitans
that results from amino acid differences between the two frogs.
As a first step in reaching a molecular understanding of the enzymatic
differences between R. sylvatica and R. clamitans
SERCA1s, we report the complete amino acid sequences for the two
proteins. Comparisons between R. sylvatica, R. clamitans,
and R. esculenta indicate that there are seven amino acids
in SERCA1 that are unique to R. sylvatica and, therefore,
may be related to the functional differences noted above (Table I).
Three of the seven amino acid substitutions (Asn546,
Ala547, and Ser643) that are unique to R. sylvatica when compared with the two other frogs are located in
the ATP binding domain. Because the ATP kinetics are different
in R. sylvatica in that Michaelis-Menten kinetics are
exhibited, it appears likely that these mutations contribute to
different enzymatic behavior exhibited by R. sylvatica
SERCA1. None of these three amino acids has been examined by
site-directed mutagenesis (12). It is possible that site-directed
mutagenesis of single amino acid residues will result in changes in
ATPase kinetics or that more than one concomitant change is required to
achieve this effect.
Three of the remaining four amino acid changes in R. sylvatica are not in regions of the SERCA1 sequence with predicted
functional roles. Gln26 is located at the beginning of an
NH2-terminal Three distinct genes code for SERCA pumps in higher vertebrates: SERCA1
(ATP2A1), SERCA2 (ATP2A2), and SERCA3
(ATP2A3) (12). More isoform diversity is brought
about by alternative processing of the SERCA primary transcripts as
follows: adult SERCA1a and neonatal SERCA1b (both expressed in
fast-twitch skeletal muscle) (39); SERCA2a (cardiac/slow-twitch
skeletal muscle) and SERCA2b (ubiquitously expressed) (12); and
SERCA3a, SERCA3b, and SERCA3c (broad cellular expression) (40). The
alternative splicing mechanism governing the processing of rabbit
SERCA1 pre-mRNA appears to be regulated developmentally . Both
isoforms present identical amino acid sequences up to amino acid 993, after which the last amino acid (Gly994) in SERCA1a is
replaced by a highly charged octapeptide stretch, DPEDERRK, in SERCA1b
(39). Previous investigations in R. esculenta (7) provided
no evidence for alternative splicing of the frog SERCA1 pre-mRNA.
One possible explanation for this lack of evidence may reside in the
use of RNA isolated from adult frog tissues (7). In contrast, we used
RNA isolated from an immediate post-metamorphic R. sylvatica
frog. This enabled us to detect the neonatal SERCA1b splice variant. In
the present report, we show that R. sylvatica SERCA1
pre-mRNA can also be spliced alternatively in the same manner as in
rabbit, i.e. the last residue (Gly994) in
SERCA1a is replaced by a charged tetrapeptide, DKPQ, in SERCA1b isoform. Remarkably, this finding demonstrates that the alternative splicing of the SERCA1 gene is conserved among amphibians and mammals.
The physiological significance of the two splice variants is not clear
yet. Moreover, COS-1 cell expression studies revealed no functional
differences between rabbit SERCA1a and SERCA1b isoforms (41).
Furthermore, it is surprising that there are no reports in the
literature indicating the expression of a SERCA2 gene in frogs, because
the expression of a unique SERCA gene resembling the SERCA2 gene was
also documented in invertebrates such as the crustacean Artemia
franciscana (42) and the insect Drosophila melanogaster
(43). Despite our efforts (Western blot analyses using anti-SERCA2
antibodies and reverse transcriptase-PCR studies using primers flanking
nonconserved nucleotide sequences among invertebrate and higher
vertebrate SERCAs), we were not able to demonstrate the existence of a
SERCA2-like gene in R. sylvatica. However, we cannot rule
out the possibility that the epitope for the anti-SERCA2 antibody is
modified in the frog protein or that the SERCA2 gene may still be
expressed but at much lower mRNA levels than SERCA1.
In conclusion, the molecular adaptation of SERCA1 in the skeletal
muscle of R. sylvatica allows this frog to exploit a unique habitat for amphibians in freshwater temporary ponds in late winter. The unusual kinetic properties of R. sylvatica SERCA1 almost
certainly result from the seven amino acid residues that are unique to
this species. Future studies will shed light on the role of these amino acids in temperature dependence, kinetics with ATP, Ca2+
cooperativity, and protein-protein interactions.
INTRODUCTION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES
EXPERIMENTAL PROCEDURES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES
) common
for both frog cDNAs, was derived from the partially determined sequence 5'-GTGCACAGATTGTGGCCAACTCCAG-3'. This primer corresponds to
the inverse complement of nucleotides 1431-1458 and 1434-1461, respectively, in R. sylvatica and R. clamitans
sequences described in this study. 5' RACE-PCR amplifications with the
adaptor primer 1/5'R
primer set were performed using the above
mentioned cycle conditions. The 5' RACE-PCR products were also
subcloned in pGEM®-T Easy vector. The composite wood and
green frog cDNA nucleotide sequences in both strands were
determined completely using several gene-specific primers and the
A.L.F.® Express DNA Sequencer (Amersham Pharmacia
Biotech). The sequencing strategy is shown in Fig. 5. For further
applications, cDNA clones containing the entire coding regions were
obtained by PCR amplifications using the following pair of primers:
fcDNA+ (5'-GGCTGAGAGGCGAGGATGGAGAACGCAC-3') corresponding to
nucleotides 189-216 and 191-219, respectively, in wood and green frog
cDNAs and fcDNA
(5'-TTATCCTTCAAGGTAGTTACGGGCAACA-3') corresponding to the inverse complement of nucleotides 3161-3188 and
3164-3191, respectively, in R. sylvatica and R. clamitans sequences. The cycle conditions were the same as above
except for a longer extension time (5 min) per cycle at 72 °C. The
amplified cDNA products (3.2 kilobase pairs) were subcloned
in either pCR®3.1 (for green frog cDNA) or
NotI cut and dephosphorylated pcDNA3 (for R. sylvatica cDNA) expression vectors (Invitrogen, San Diego, CA.
The accuracy of the nucleotide sequence was verified by complete sequencing of several clones for each frog SERCA cDNA.
(5'-GCTGCATTCCTTTTACTGTGGTTTAT-3') corresponding to the inverse
complement of nucleotides 3225-3250 in wood frog SERCA cDNA. The
adaptor-ligated ds-cDNA served as template DNA. PCR amplification
was carried out for 35 cycles, each cycle consisting of 30 s at
94 °C, 30 s at 55 °C, and 30 s at 72 °C. The PCR
products were subcloned in the pGEM®-T Easy vector, and
several positive clones were sequenced.
RESULTS
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES
View larger version (29K):
[in a new window]
Fig. 1.
SDS-polyacrylamide gel analysis of R. sylvatica and R. clamitans SR.
A, Coomassie Blue staining of SDS-7.5% polyacrylamide gel.
Positions of molecular mass standards and probable SERCA proteins are
indicated. B, results of immunoblotting 1-µg samples of
wood frog (R. sylvatica) and green frog (R.
clamitans) with monoclonal anti-SERCA1 followed by
chemiluminescent detection. Antibodies against SERCA2 and SERCA3 did
not label any frog proteins (data not shown).
View larger version (11K):
[in a new window]
Fig. 2.
Temperature dependence of R. sylvatica and R. clamitans SR
Ca2+-ATPase and Ca2+ transport
activities. SR membranes were assayed as described under
"Experimental Procedures." The specific activity of the green frog
SR preparation was 2.7 µmol of ATP hydrolyzed/min/mg at 37 °C,
whereas the specific activity of the wood frog SR preparations was 0.8 in A and 3.0 in B. A, temperature
dependence of Ca2+-ATPase activity of wood and green frogs
with 10 mM MgATP and 10 µM Ca2+
plotted in the Arrhenius format. Error bars indicate the
mean ± S.D. from the mean for triplicate assays at each
temperature. *, indicates significant differences at the
p < 0.005 level with a paired t test.
B, temperature dependence of Ca2+ transport
activity of wood and green frog SR plotted in the Arrhenius format.
Each point is the average of two determinations at 5 mM
MgATP and 5 µM Ca2+.
3 1/K)
are significantly different at the p < 0.005 level
t test. Calculation of the activation energies by linear
fitting below 20 °C yields a significantly higher value for the
green frog (28 ± 1.8 kcal/mol) (mean ± S.E.) than for the
wood frog (23.7 ± 0.6). Extrapolation to 0 °C indicates that
the wood frog will exhibit a 2.1-fold higher activity than the green
frog. Thus, although the activity is low, the wood frog exhibits
significantly higher activity.
View larger version (16K):
[in a new window]
Fig. 3.
Kinetics of wood and green frog SR as a
function of MgATP concentration. SR membranes were assayed as
described under "Experimental Procedures" with 10 µM
Ca2+ and the indicated concentrations of MgATP. Each assay
was done in quadruplicate, and the points and error
bars indicate the mean ± S.D. Curves are fitted to a
Michaelis-Menten model for wood frog and a two-site model for green
frog with a Km of 5.0 µM and a
Vmax of 0.49 for the first site and a
Km of 0.34 mM and
Vmax of 0.34 for the second site. The wood frog
exhibits a Km = 0.12 ± 0.03 mM
(mean ± S.D.), whereas the green frog has a
K1/2 = 0.07 mM.
View larger version (15K):
[in a new window]
Fig. 4.
Kinetics of wood and green frog SR as a
function of free Ca2+ concentration. SR membranes were
assayed as described under "Experimental Procedures" with 10 mM MgATP and the indicated concentrations of free
Ca2+. Each assay was done in quadruplicate, and the
points and error bars indicate the mean ± S.D. The y axis shows the reaction velocity relative to the
maximal velocity at 0.1 mM Ca2+
(v/Vmax). Fitting the data to a
positive cooperativity model yielded Hill coefficients of 0.9 ± 0.2 (± S.D.) for the wood frog and 1.5 ± 0.2 (± S.D.) for the
green frog.
) was generated to amplify the 5'
cDNA ends. The resulting PCR products (1.5 kilobase pairs)
were subcloned and sequenced. Using several other sequencing primers,
the 3' and 5' overlapping PCR products were completely sequenced in
both strands. The composite nucleotide sequences of R. sylvatica and R. clamitans SERCA cDNAs were
determined (data not shown) (GenBankTM accession numbers
AJ298901 and AJ298902). The sequencing strategy is shown in Fig.
5. Wood and green frog SERCA cDNAs
are, respectively, 3790 and 3793 nucleotides in length. The nucleotide difference is due to the presence of three more nucleotides in the
5'-untranslated region of R. clamitans cDNA (206 versus 203 nucleotides in the wood frog sequence). The
nucleotide sequence (AGGATGG) surrounding the translation
initiation codon (ATG) is found in an acceptable context
characteristic of initiation of eukaryotic translation (25). Two
adjacent consensus polyadenylation signals (ATTAAA) (26) in the
605-nucleotide 3'-untranslated region of both frog SERCA cDNAs are
located, respectively, 18 and 12 nucleotides upstream of the beginning
of the poly(A) tract.
View larger version (12K):
[in a new window]
Fig. 5.
Sequencing strategy for R. sylvatica and R. clamitans SERCA1a cDNA
clones. The full-length composite clone shown below the
scale line is representative of either frog SERCA1a
cDNA. A black arrow represents the open reading frame,
and the gray flanking lines denote the untranslated
sequences. The oligonucleotides used for sequencing the cDNA clones
are illustrated by arrowheads above and beneath the cDNA
clone. The primers PH+ and 5'R used for the RACE-PCR experiments are
indicated as well. The resulting overlapping 5' and 3' RACE-PCR
products are shown at the bottom as heavy lines
with arrows indicating the extent and direction of
amplification.
View larger version (57K):
[in a new window]
Fig. 6.
Amino acid similarity comparison. The
deduced amino acid sequence of R. sylvatica SERCA1a cDNA
(RS SERCA1a) is compared with those of R. clamitans (RC SERCA1a) from this study, R. esculenta (RE SERCA1a) (2), and rabbit (RAB
SERCA1a) (10). Residues of the green frog, European frog, and
rabbit SERCA1a that are identical with the corresponding residues of
wood frog SERCA1a are represented by dashes. The amino acid
numbers for wood frog SERCA1a are shown at the right. The
hydrophobic putative transmembrane domains (M1-M10), phosphorylation
site (Phos Site), the fluorescein isothiocyanate
(FITC) binding site (7), and the FSBA/CIRATP
(FSBA/CIRATP) binding region (8, 9) are labeled and
indicated by bars above the sequence. Positions
of the amino acid residues unique to R. sylvatica enzyme are
indicated by arrowheads.
Unique amino acid substitutions in R. sylvatica
as primers and
the adaptor-ligated ds-cDNA as template. After amplification, two
products of 159 and 117 base pairs were detected by 1.8% agarose gel
electrophoresis (data not shown). Subcloning and subsequent sequencing
confirmed that the 159-base pair PCR fragment corresponded to SERCA1a,
whereas the second smaller fragment represented a novel splice variant,
SERCA1b. The retention or exclusion of a 42-base pair optional
exon (Fig. 7) gives rise, respectively, to SERCA1a and SERCA1b splice variants. The insertion of the optional exon in wood frog SERCA1a occurs immediately after nucleotide 2980 (relative to the ATG codon), which, interestingly, represents the point
of divergence between the different splice variants in the related
SERCA1, SERCA2, and SERCA3 genes. In turn, the nucleotide sequence
divergence in SERCA1a splice variants is reflected at the protein level
by a generation of SERCA1 isoforms with different COOH termini. Both
isoforms present identical amino acid sequences up to amino acid 993 (Glu), after which the last residue (Gly994) in SERCA1a is
replaced in SERCA1b by a tetrapeptide stretch (DKPQ997). The significance of the different COOH
termini has still to be investigated.
View larger version (21K):
[in a new window]
Fig. 7.
Alternative splicing of the SERCA1
pre-mRNA in R. sylvatica. The nucleotide
sequences of the PCR fragments corresponding to SERCA1a and SERCA1b
splice variants are compared. Nucleotides are numbered at
the right of the sequence. The PCR primers used
(SPLICE+ and SPLICE ) are indicated by
arrows above their corresponding sequence. Below
the nucleotide sequence, the translation products representing the
COOH-terminal parts of SERCA1a and SERCA1b are compared (amino acids
are numbered at the right). The optional exon (42 nucleotides) flanked by two slashes is shown in
bold. The stop codon, TAA, is indicated by three consecutive
asterisks.
DISCUSSION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES
-helix that has no known functional role.
Leu838 is located in the M7 transmembrane segment that is
thought to be located on the outer surface of the proposed
Ca2+ channel and to have no direct role in Ca2+
transport (15). Ser877 is located in the extracellular loop
between M7 and M8. Mutagenesis of this residue has no effect on
activity (12). Ser650 (Thr650 in R. sylvatica) was originally proposed to be in the nucleotide binding
domain (14), but the recent determination of the structure of SERCA1 at
2.6-Å resolution indicates that this residue is in the phosphorylation
domain (15). This position has not been examined by site-directed mutagenesis.
![]() |
ACKNOWLEDGEMENTS |
---|
We thank Timothy D. Cummins for expert technical assistance. We also thank Dr. Jens Peter Andersen (Department of Physiology, University of Aarhus, Denmark) for the gift of the rabbit SERCA1a cDNA.
![]() |
FOOTNOTES |
---|
* This work was supported in part by the Special Research Infrastructure Fund, University of Louisville (to W. L. D.) and the University Poles of Attraction Programme IUAP P4/23, Belgian State, Prime Minister's Office, Federal Office for Scientific, Technical, and Cultural Affairs (to F. W.).The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.
The nucleotide sequence(s) reported in this paper has been submitted to the GenBankTM/EMBL Data Bank with accession number(s) AJ298901 for R. sylvatica and AJ298902 for R. clamitans SERCA1s.
§ Both authors contributed equally to this work.
¶ Present address: The Wellcome Trust Center for Human Genetics, University of Oxford, Roosevelt Dr., Oxford OX3 7BN, United Kingdom.
** To whom correspondence should be addressed: Dept. of Biochemistry and Molecular Biology, School of Medicine, University of Louisville, Louisville, KY 40292. Tel.: 502-852-5227; Fax: 502-852-6222; E-mail: bill.dean@louisville.edu.
Published, JBC Papers in Press, October 23, 2000, DOI 10.1074/jbc.M007719200
![]() |
ABBREVIATIONS |
---|
The abbreviations used are: SERCA, sarco(endo)plasmic reticulum Ca2+-ATPase; PCR, polymerase chain reaction; RACE, rapid amplification of cDNA ends; SR, sarcoplasmic reticulum; Tes, 2-{[2-hydroxy-1,1-bis(hydroxymethyl)ethyl]amino}ethanesulfonic acid; ds, double-stranded; FSBA, 5'-p-fluorosulfonylbenzoyladenosine.
![]() |
REFERENCES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
1. |
Storey, K. B.,
and Storey, J. M.
(1988)
Physiol. Rev.
68,
27-84 |
2. | Mommsen, T. P., and Storey, K. B. (1992) Gen. Comp. Endocrinol. 87, 44-53[Medline] [Order article via Infotrieve] |
3. |
Constanzo, J. P.,
Lee, R. E.,
and Lortz, P. H.
(1993)
Am. J. Physiol.
265,
R721-R725 |
4. |
Dalo, N. L.,
Hackman, J. C.,
Storey, K. B.,
and Davidoff, R. A.
(1995)
J. Exp. Biol.
198,
1765-1774 |
5. | Hemmings, S. J., and Storey, K. B. (1996) Cell Biochem. Funct. 14, 139-148[CrossRef][Medline] [Order article via Infotrieve] |
6. | Hemmings, S. J., and Storey, K. B. (1996) Can. J. Physiol. Pharmacol. 72, 1552-1560 |
7. | Vilsen, B., and Andersen, J. P. (1992) FEBS Lett. 306, 213-218[CrossRef][Medline] [Order article via Infotrieve] |
8. | Ogawa, Y. (1970) J. Biochem. (Tokyo) 67, 667-683[Medline] [Order article via Infotrieve] |
9. | Eletr, S., and Inesi, G. (1972) Biochim. Biophys. Acta 290, 178-185[Medline] [Order article via Infotrieve] |
10. | Kossler, F., and Kuchler, G. (1987) Biomed. Biochim. Acta 46, 815-822[Medline] [Order article via Infotrieve] |
11. |
Johnson, J. D.,
Jiang, Y.,
and Flynn, M.
(1997)
Am. J. Physiol.
272,
C1437-C1444 |
12. |
MacLennan, D. H.,
Rice, W. J.,
and Green, N. M.
(1997)
J. Biol. Chem.
272,
28815-28818 |
13. |
MacLennan, D. H.
(1970)
J. Biol. Chem.
245,
4508-4518 |
14. | MacLennan, D. H., Brandl, C. J., Korczak, B., and Green, N. M. (1985) Nature 316, 696-700[Medline] [Order article via Infotrieve] |
15. | Toyoshima, C., Nakasako, M., Nomura, H., and Ogawa, H. (2000) Nature 405, 647-656[CrossRef][Medline] [Order article via Infotrieve] |
16. | Bull, R., Marengo, J. J., Suarez-Isla, B. A., Donoso, P., Sutko, J. L., and Hidalgo, C. (1989) Biophys. J. 56, 749-756[Abstract] |
17. | Laemmli, U. K. (1970) Nature 227, 680-685[Medline] [Order article via Infotrieve] |
18. | Dean, W. L., and Quinton, T. M. (1995) Cell Calcium 17, 65-70[Medline] [Order article via Infotrieve] |
19. | Dean, W. L., and Tanford, C. (1978) Biochemistry 17, 1683-1690[Medline] [Order article via Infotrieve] |
20. | Quinton, T. M., and Dean, W. L. (1992) Biochem. Biophys. Res. Commun. 184, 893-899[Medline] [Order article via Infotrieve] |
21. | Chirgwin, J. M., Przybla, A. E., MacDonald, R. J., and Rutter, W. J. (1979) Biochemistry 18, 5294-5299[Medline] [Order article via Infotrieve] |
22. | Neet, E. E., and Green, N. M. (1977) Arch. Biochem. Biophys. 178, 588-597[Medline] [Order article via Infotrieve] |
23. |
Verjovski-Almeida, S.,
and Silva, J. L.
(1981)
J. Biol. Chem.
256,
2940-2944 |
24. | Frohman, M. A. (1990) in PCR Protocols: A Guide to Methods and Applications (Innis, M. A. , Gelfand, D. H. , Sninsky, J. J. , and White, T. J., eds) , pp. 28-38, Academic Press, San Diego |
25. | Kozak, M. (1987) Nucleic Acids Res. 15, 8125-8148[Abstract] |
26. | Proudfoot, N. J., and Brownlee, G. G. (1976) Nature 263, 211-214[Medline] [Order article via Infotrieve] |
27. | Mitchinson, C., Wilderspin, A. F., Trinnaman, B. J., and Green, N. M. (1982) FEBS Lett. 146, 87-92[CrossRef][Medline] [Order article via Infotrieve] |
28. | Ohta, T., Nagano, K., and Koshida, M. (1986) Proc. Natl. Acad. Sci. U. S. A. 83, 2071-2075[Abstract] |
29. | Ovchinnikov, Y. A., Dzhandzugazyan, K. N., Lutsenko, S. V., Mustayev, A. A., and Modyanov, N. N. (1987) FEBS Lett. 217, 111-116[CrossRef][Medline] [Order article via Infotrieve] |
30. | Martonosi, A. N. (1996) Biochim. Biophys. Acta 1275, 111-117[Medline] [Order article via Infotrieve] |
31. |
Dean, W. L.,
and Gray, R. D.
(1980)
J. Biol. Chem.
255,
7514-7516 |
32. |
Deamer, D. W.
(1973)
J. Biol. Chem.
248,
5477-5485 |
33. | Inesi, G., Millman, M., and Eletr, S. (1973) J. Mol. Biol. 81, 483-504[Medline] [Order article via Infotrieve] |
34. | Lee, A. G., Birdsall, N. J. M., Metcalfe, J. C., Toon, P. A., and Warren, G. B. (1974) Biochemistry 13, 3699-3705[Medline] [Order article via Infotrieve] |
35. | Hesketh, T. R., Smith, G. A., Houslay, M. D., McGill, K. A., Birdsall, N. J. M., Metcalfe, J. C., and Warren, G. B. (1976) Biochemistry 15, 4145-4151[Medline] [Order article via Infotrieve] |
36. |
Kalabokis, V.,
and Hardwicke, P.
(1988)
J. Biol. Chem.
263,
15184-15188 |
37. | Chini, E. N., de Toledo, F. G. S., Albuquerque, M. C., and de Meis, L. (1993) Biochem. J. 293, 469-473[Medline] [Order article via Infotrieve] |
38. | Vrbjar, N., Simatos, G. A., and Keough, K. M. W. (1990) Biochim. Biophys. Acta 1030, 94-100[Medline] [Order article via Infotrieve] |
39. |
Korczak, B.,
Zarain-Herzberg, A.,
Brandl, C. J.,
Ingles, C. J.,
Green, N. M.,
and MacLennan, D. H.
(1988)
J. Biol. Chem.
263,
4813-4819 |
40. |
Dode, L.,
De Greef, C.,
Mountian, I.,
Attard, M.,
Town, M. M.,
Casteels, R.,
and Wuytack, F.
(1998)
J. Biol. Chem.
273,
13982-13994 |
41. | Maruyama, K., and MacLennan, D. H. (1988) Proc. Natl. Acad. Sci. U. S. A. 85, 3314-3318[Abstract] |
42. |
Escalante, R.,
and Sastre, L.
(1994)
J. Biol. Chem.
269,
13005-13012 |
43. | Magyar, A., Bakos, É., and Vàradi, A. (1995) Biochem. J. 310, 757-763[Medline] [Order article via Infotrieve] |