Sie sind auf Seite 1von 6

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 276, No. 50, Issue of December 14, pp.

47094 –47099, 2001


© 2001 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Phe71 Is Essential for Chaperone-like Function in ␣A-crystallin*


Received for publication, August 13, 2001, and in revised form, September 21, 2001
Published, JBC Papers in Press, October 11, 2001, DOI 10.1074/jbc.M107737200

Puttur Santhoshkumar‡ and K. Krishna Sharma‡§¶


From the Departments of ‡Ophthalmology and §Biochemistry, University of Missouri, Columbia, Missouri 65212

Experiments with mini-␣A-crystallin (KFVIFLD- the lens (2, 19 –21). During chaperone-like action, hydrophobic
VKHFSPEDLTVK) showed that Phe71 in ␣A-crystallin surfaces in ␣-crystallin interact with specific sites in non-na-
could be essential for the chaperone-like action of the tive target proteins (22–24). Earlier we were able to map the
protein (Sharma, K. K., Kumar, R. S., Kumar, G. S., and site in ␣A- and ␣B-crystallin responsible for chaperone-like
Quinn, P. T. (2000) J. Biol. Chem. 275, 3767–3771). In the action using photoactive cross-linkers and hydrophobic probes
present study we replaced Phe71 in rat ␣A-crystallin (25–27). Our studies with bis-ANS1 and the hydrophobic pro-
with Gly by site-directed mutagenesis and then com- tein mellitin have shown that there is an overlapping of chap-
pared the structural and functional properties of the erone site and hydrophobic site in ␣A-crystallin. Further, using
mutant protein with the wild-type protein. There were a synthetic peptide (mini-␣A-crystallin), we were able to dem-
no differences in molecular size or intrinsic tryptophan onstrate the importance of sequence 70 – 88 in the chaperone-
fluorescence between the proteins. However, 1,1ⴕ-bi(4-
like action of ␣A-crystallin (28). The experiments with trun-
anilino)naphthalene-5,5ⴕ-disulfonic acid interaction in-

Downloaded from www.jbc.org by guest, on October 28, 2010


cated forms of mini-␣A-crystallin had suggested that Phe71 in
dicated a higher hydrophobicity for the mutant protein.
␣A-crystallin might be critical for chaperone-like function. In
Both wild-type and mutant proteins displayed similar
secondary structure during far UV CD experiments. the present study, we did a site-directed mutagenesis of Phe71
Near UV CD signal showed a slight difference in the to Gly in ␣A-crystallin and compared the structural and func-
tertiary structure around the 285–295 region for the two tional properties of this mutant protein with the wild-type
proteins. The mutant protein was totally inactive in sup- protein.
pressing the aggregation of reduced insulin, heat-dena- Recombinant ␣A- and ␣B-crystallins show similar structural
tured citrate synthase, and alcohol dehydrogenase. and functional properties to crystallins isolated from lens tis-
However, a marginal suppression of ␤L-crystallin aggre- sues and are widely used in the characterization of the protein.
gation was observed when mutant ␣A-crystallin was in- Site-directed mutations of recombinant protein provide an ex-
cluded. These results suggest that Phe71 contributes to cellent means of studying the role of constituent amino acids in
the chaperone-like action of ␣A-crystallin. Therefore we the functional properties of the protein. Earlier, several site-
conclude that the 70 – 88-region in ␣A-crystallin, identi- directed mutations were conducted on ␣-crystallin either to
fied by us earlier, is the functional chaperone site in identify the region responsible for chaperone-like function or to
␣A-crystallin. explain the role of ␣-crystallin in hereditary cataracts and
certain other diseases (29 –38). The majority of these studies
report either no change in chaperone-like function or a partial
␣-Crystallins are major refractive proteins in the vertebrate loss of this function. We report here, for the first time, a
eye lens. When isolated from the lens they exist as polydisperse complete loss in the functional property of a mutant ␣A-crys-
aggregates having an average molecular mass of ⬇800 kDa (1, tallin at and slightly above physiological temperatures. The
2). ␣-Crystallin is composed of two subunits, ␣A and ␣B, which results also indicate the presence of additional sites in ␣-crys-
have considerable sequence homology between them and with tallin that become available at elevated temperatures. We con-
other heat shock proteins (3, 4). Recently ␣-crystallin subunits clude here that the region identified by us earlier as chaperone
were also reported to be present in nonlenticular tissues like site contributes to the chaperone-like activity of ␣A-crystallin.
heart, brain, and kidney (5– 8). The significance of their pres-
ence in nonlenticular tissues is not clear. However, the in- EXPERIMENTAL PROCEDURES
creased expression of ␣B-crystallin observed in a variety of Preparation of the Mutant Clone—Rat ␣A-crystallin cDNA cloned in
neurological disorders has drawn significant medical attention pET21b was kindly donated by Dr. Suraj Bhat (UCLA). ␣AF71G mu-
tant was constructed using a QuikChange site-directed mutagenesis kit
(9 –11). Like other small heat shock proteins, ␣-crystallin can
(Stratagene). The following set of primers were used: 5⬘-CTGACCGG-
sequester certain unfolding proteins in vitro, by preventing GACAAGGGTGTCATCTTCTTGG-3⬘ and 5⬘-CCAAGAAGATGACAC-
their aggregation and insolubilization (12–15). Both subunits CCTTGTCCCGGTCAG-3⬘. The mutation was confirmed by automated
of ␣-crystallin show chaperone-like activity to different extents DNA sequencing.
(16 –18). Complex formation with ␤L- and ␥-crystallin and de- Expression and Purification of Wild-type and Mutant ␣A-crystallin—
creased chaperone-like function during aging has indicated the The proteins were expressed in Escherichia coli BL21(DE3) cells (No-
vagen) as described by Horwitz et al. (39). The proteins were isolated
importance of ␣-crystallin in maintaining the transparency of
from the cell pellet using Bugbuster protein extract reagent (Novagen).
In brief, 1 g of cells was suspended in 5 ml of reagent at room temper-
* This work is supported in part by National Institutes of Health ature and vortexed gently. Protease inhibitor mixture set III (Novagen)
Grant EY11981 and a grant-in-aid from Research to Prevent Blindness. was then added. The cell suspension was treated with 1 ␮l (25 units) of
The costs of publication of this article were defrayed in part by the benzonase/ml of Bugbuster reagent and incubated at room temperature
payment of page charges. This article must therefore be hereby marked on a shaking platform for 30 min. The extract was centrifuged at
“advertisement” in accordance with 18 U.S.C. Section 1734 solely to
indicate this fact.
1
¶ To whom correspondence should be addressed: Mason Eye Inst., The abbreviations used are: bis-ANS, 1,1⬘-bi(4-anilino)naphthalene-
Dept. of Ophthalmology, 1 Hospital Dr., University of Missouri, Colum- 5,5⬘-disulfonic acid; CS, citrate synthase; ADH, alcohol dehydrogenase;
bia, MO 65212. E-mail: Sharmak@health.missouri.edu. HPLC, high pressure liquid chromatography.

47094 This paper is available on line at http://www.jbc.org


F71G Mutant of ␣A-crystallin 47095
17,000 ⫻ g for 2 h, and the supernatant was filtered through a 0.2-␮m
filter. The filtrate was loaded onto a Bio-Rad High Q anion exchange
column and eluted using a linear gradient of NaCl (0 –1 M) in 20 mM
Tris-HCl (pH 8) at a flow rate of 2 ml/min. The fractions containing the
recombinant crystallin (determined by SDS-PAGE) were pooled and
concentrated. The concentrated protein was purified further on a C18
reverse phase HPLC using a water-acetonitrile gradient containing
0.1% trifluoroacetic acid. The peaks corresponding to wild-type and
mutant ␣A-crystallin were pooled, dried on a Speedvac, resuspended in
6 M urea, and dialyzed extensively against 0.05 M phosphate (PO4)
buffer containing 0.15 M NaCl, with several changes, for a period of 2
days. The purity of the proteins was checked by SDS-PAGE, and the
mass was determined by mass spectrometry.
Molecular Size Determination—Size determination was carried out
using a Amersham Biosciences Hiload 16/60 Superdex 200 gel filtration
column equilibrated with 0.05 M PO4 buffer containing 0.15 M NaCl (pH
7.4). The mass was calculated from the calibration curve generated by FIG. 1. SDS-PAGE analysis of recombinant protein purifica-
using Sigma molecular weight marker standards. tion. Lane 1, crude cell extract; lane 2, recombinant protein after ion
Tryptophan Fluorescence Measurements—The intrinsic fluorescence exchange; lane 3, recombinant protein after reverse phase; lane 4,
spectra of the wild-type and mutant ␣A-crystallin were recorded using molecular mass markers.
a Jasco spectrofluorometer FP-750. Protein samples of 200 ␮g/ml in
0.05 M PO4 buffer containing 0.15 M NaCl were used. The excitation was
set to 295 nm, and the emission was recorded between 300 and 400 nm.
bis-ANS Fluorescence Measurement—To 100 ␮g of wild-type and
mutant protein taken in 0.05 M PO4 containing 0.15 M NaCl (pH 7.4)
was added 20 ␮l of 10 mM bis-ANS dissolved in ethanol. The sample was

Downloaded from www.jbc.org by guest, on October 28, 2010


excited at 365 nm, and the fluorescence spectrum was measured be-
tween 400 and 600 nm using a Jasco spectrofluorometer.
Circular Dichroism Studies—Protein secondary and tertiary struc-
tural changes were investigated by far and near UV CD measurements
using an AVIV circular dichroism spectrometer. The concentration of
the proteins used was 1.5 and 0.35 mg/ml for near and far UV CD,
respectively. The reported CD spectra are the averages of four scans.
Chaperone-like Activity—The ability of the wild-type and mutant
proteins to prevent protein aggregation was determined using several
substrates. The extent of aggregation was measured by monitoring the
light scattering at 360 nm in a Shimadzu spectrophotometer.
Insulin Aggregation Assay—The aggregation of insulin (0.4 mg/ml)
(Sigma) in 0.05 M PO4 buffer containing 0.15 M NaCl (pH 7.2) was
initialized by the addition of 25 ␮l of 1 M dithiothreitol in the presence
of wild-type and mutant proteins. The aggregation was monitored at
room temperature.
CS Aggregation Assay—CS (75 ␮g) (Roche Molecular Biochemicals)
in 1 ml of 40 mM HEPES-NaOH buffer (pH 7.4) was heated at 43 °C for
1 h in the presence of various amounts of mutant and wild-type pro-
teins. The light scattering was measured as described above.
ADH Aggregation Assay—ADH (250 ␮g) (Sigma) was heated at 45 °C
in the presence of various amounts of mutant and wild-type ␣A-crys-
tallin in 0.05 M PO4 buffer containing 0.15 M NaCl (pH 7.4). The
scattering of light was measured up to 100 min.
␤L-Crystallin Aggregation Assay—Purified bovine ␤L-crystallin (250
␮g) in 0.05 M PO4 buffer containing 0.15 M NaCl (pH 7.4) was heated at
55 °C in the presence of different amounts of recombinant proteins. The FIG. 2. Gel permeation profiles of refolded wild-type and mu-
aggregation was monitored up to 1 h. tant protein on a Superdex 200 column.

RESULTS also observed a similar mass for reconstituted homopolymers of


Characterization of Recombinant ␣A-crystallin—In the pres- bovine lens ␣A-crystallin.
ent study, the recombinant proteins were purified by a combi- The structural differences between the wild-type and mutant
nation of ion exchange and reverse phase HPLC. The purified proteins were analyzed by spectroscopic methods. Tryptophans
proteins were dissolved in urea and refolded by extensive dial- of the protein have a fixed solvent accessibility, and any change
ysis. The proteins thus obtained were highly pure (Fig. 1). The in their environment leads to an altered fluorescence emission
ES mass spectrometry analysis revealed molecular masses of pattern and intensity. Our results show no change in the tryp-
19,799 and 19,709 daltons, which would be expected for the tophan region of the recombinant proteins, as evidenced by the
wild-type and mutant ␣A-crystallin respectively. Like the similar fluorescence emission maximum (340 nm) and intensity
␣-crystallin subunits isolated from eye lens, recombinant pro- (Fig. 3). Unfolding of proteins increases the exposure of hydro-
teins exist in oligomeric form. For analysis of the molecular phobic surfaces that can be probed with bis-ANS fluorescence
mass of the homoaggregates, the purified recombinant proteins (26). We see an increase in bis-ANS binding to the ␣AF71G
were chromatographed on a Superdex-200 column. Both wild- mutant (Fig. 4), indicating an increased hydrophobicity com-
type and ␣AF71G mutant proteins showed similar elution pro- pared with wild-type protein.
files, corresponding to an oligomeric mass of 7.1 ⫻ 105 daltons The secondary and tertiary structures of wild-type and mu-
(Fig. 2). This is slightly higher than the earlier published tant ␣A-crystallin were determined by far and near UV CD
values for rat ␣A-crystallin (37). The discrepancy can be attrib- spectral analysis. The far UV profile showed a characteristic
uted to the different buffer conditions used in analysis, because ␤-sheet conformation with a slight increase in the negative
it is known that the mass of the purified protein varies depend- intensity of the mutant protein (Fig. 5). Both proteins showed
ing on the buffer condition (40, 41). During these studies we similar amounts of ␣-helix, ␤-sheet, and random coil (42). Near
47096 F71G Mutant of ␣A-crystallin

Downloaded from www.jbc.org by guest, on October 28, 2010


FIG. 3 Intrinsic fluorescence intensity of purified wild-type
and mutant ␣A-crystallin. Protein samples (200 ␮g) in phosphate FIG. 5 Far UV CD spectra of recombinant proteins.
buffer were excited at 295 nm.

FIG. 6. Near UV CD spectra of recombinant proteins.


FIG. 4. Interaction of bis-ANS with mutant and wild-type
␣A-crystallin.
of insulin results in the separation of the subunits and precip-
UV CD spectra showed a slight increase in the negative inten- itation of B chain that can be followed by measurement of light
sity of the mutant protein (Fig. 6). Although significant por- scattering. The presence of ␣-crystallin subunits in the assay
tions of the near UV spectra for the two proteins were similar, prevents the aggregation of insulin B chain, and the solution
only minor changes were seen in the 285–295 nm region of the remains clear. Fig. 7 shows the dithiothreitol-induced aggrega-
spectra, suggesting some differences in the tyrosine and/or tion kinetics of insulin in the presence of both wild-type and
tryptophan microenvironments of the mutant protein com- mutant ␣A-crystallins. The wild-type protein showed suppres-
pared with the wild-type ␣A-crystallin. Surprisingly enough, sion of insulin B chain aggregation that increased with the
there was no alteration in the signal caused by phenylalanine concentration of the protein in the assay tube. However, the
in the 250 –270-nm region. In summary, the data in Fig. 6 do mutant ␣A-crystallin completely failed to prevent the forma-
not suggest a significant difference in the tertiary structure tion of light-scattering aggregates. In fact, a marginal increase
between wild-type and mutant ␣A-crystallin. in light scattering was observed in some assays. Higher con-
The Chaperone-like Activity of F71G ␣A-crystallin—The con- centrations of mutant protein had no effect on the aggregation
sequence of mutation on recombinant crystallin chaperone-like of polypeptide. The chaperone-like activity of the recombinant
activity was determined under different conditions. Reduction proteins was also investigated at different temperatures. Fig. 8
F71G Mutant of ␣A-crystallin 47097

Downloaded from www.jbc.org by guest, on October 28, 2010


FIG. 7 Reduction of insulin (0.4 mg) by dithiothreitol in the
presence of wild-type and mutant ␣A-crystallin. A, insulin alone;
B–D, insulin with 0.1, 0.2, and 0.3 mg of mutant ␣A-crystallin, respec-
tively; E and F, insulin with 0.05 and 0.1 mg of wild-type ␣A-crystallin,
respectively.

FIG. 9. Thermal aggregation of ADH (250 ␮g) in the presence of


various amounts of wild-type and mutant ␣A-crystallin at 45 °C.
A, in the presence of wild type; B, in the presence of mutant
␣A-crystallin.

at 80 min was comparable with ADH by itself (Fig. 9B). In-


creasing the concentration of mutant protein had no effect on
the aggregation of ADH. We also compared the abilities of
mutant and wild-type ␣A-crystallin to prevent the heat-in-
duced aggregation of ␤L-crystallin at 55 °C (Fig. 10). Unlike
other substrates, mutant ␣A-crystallin showed a significant
protection of ␤L-crystallin with increasing concentration. How-
ever, compared with the wild-type ␣A-crystallin, the mutant
␣A-crystallin was 6 –10-fold less effective in suppressing
␤L-aggregation.
DISCUSSION
␣A-crystallin subunit has been categorized into three do-
mains: an N-terminal domain containing residues 1– 66, a C-
terminal or ␣-crystallin domain (central core) comprising resi-
FIG. 8. Aggregation of CS at 43 °C in the presence of wild-type dues 64 –105, and an extended C-terminal including residues
and mutant ␣A-crystallin. A, CS (75 ␮g); B, CS ⫹ mutant (50 ␮g); C, 106 –173 (43– 45). Most of the mutational studies on ␣A-crys-
CS ⫹ mutant (100 ␮g); D, CS ⫹ wild type (25 ␮g); E, CS ⫹ wild type (50 tallin were conducted either on the N-terminal domain or the
␮g).
C-terminal extension. Derham and Harding (46) have reviewed
the mutations conducted by different laboratories on ␣-crystal-
shows the thermal aggregation of CS in the presence of wild- lin and have recently reanalyzed the chaperone-like activity of
type and mutant proteins. Although the wild-type protein (50 several mutants (35). In the present study, we produced an
␮g) completely suppressed the aggregation of CS (75 ␮g), the ␣A-crystallin mutant by substituting Phe71 in the core region
mutant protein, as with insulin, failed to prevent the aggrega- with a neutral amino acid Gly. This residue is highly conserved
tion of denaturing CS. We also analyzed the ability of recom- in ␣A-crystallin and is located in the region identified as the
binant proteins to suppress the aggregation of ADH at 45 °C. chaperone site of ␣A-crystallin (28). Biophysical characteriza-
The wild-type ␣A-crystallin showed increased suppression of tion of the recombinant protein revealed no change in the
denaturing protein aggregation with increasing concentration oligomer size or tryptophan fluorescence. Because the ␣-crys-
(Fig. 9A). Although the mutant protein appeared to suppress tallin molecule has only one tryptophan at position 9, the
the aggregation of ADH at initial time points, the aggregation intrinsic tryptophan fluorescence data may be of limited value
47098 F71G Mutant of ␣A-crystallin
fluorescence intensity, it is possible that the alteration is in the
tyrosine region. Interestingly, two tyrosine residues in ␣A-
crystallin are found near the bis-ANS-binding region 50 –54
(27). This may explain the increased bis-ANS binding of the
mutant protein. However, it is unlikely that such a minor
difference in the near UV CD signal would completely abolish
the chaperone-like activity of the molecule. The ␣AR116C mu-
tant, with structural alterations at many levels, showed only a
25% decrease in chaperone-like activity (37, 38). Further, it has
been shown that ␣-crystallin could preserve its chaperone func-
tion despite some irreversible structural changes (50).
We have measured the chaperone-like function of the
␣AF71G mutant under different conditions and observed a
complete loss in the activity of the mutant up to 45 °C. How-
ever, at elevated temperatures the mutant showed some sup-
pression of ␤L-crystallin aggregation. It has been shown that
␣-crystallin undergoes a structural transition around 55 °C,
resulting in the exposure of more hydrophobic patches (23, 24,
26). Our study on the stabilization of restriction enzyme (51) as
well as studies conducted by others (35, 52) indicates the pres-
ence of multiple sites in ␣-crystallin for chaperone function.
Based on the experiments with mini-␣A-crystallin (28) and the

Downloaded from www.jbc.org by guest, on October 28, 2010


FIG. 10. Aggregation of ␤L-crystallin (250 ␮g) in the presence of
recombinant proteins. A, ␤L-crystallin; B, ␤L-crystallin ⫹ mutant ␣A complete loss of chaperone-like function of the mutant protein
(100 ␮g); C, ␤L-crystallin ⫹ mutant ␣A (200 ␮g); D, ␤L-crystallin ⫹ at physiological temperatures in this study, we conclude that
mutant ␣A (300 ␮g); E, ␤L-crystallin ⫹ wild type ␣A (50 ␮g); F, ␤L- the region identified by us earlier (residues 71– 88) contributes
crystallin ⫹ wild type ␣A (100 ␮g).
to the chaperone-like function.
Plater et al. (30) reported that the F27R mutation in the
to describe the structural changes in the central core or C- N-terminal domain of ␣B-crystallin completely abolishes its
terminal domain as a consequence of the mutation. However, it chaperone-like activity at higher temperatures, which led them
will be a valuable tool for analyzing the stability of the N- to conclude that this conserved residue is vital for chaperone
terminal domain of the mutant protein. When the mutant function. Their report is controversial because later studies
protein was heated up to 60 °C, we did not observe any aggre- have shown that the mutant F27R is fully active (34, 35).
gation or shift in tryptophan fluorescence emission wavelength Earlier, work showed that proteins resulting from mutation of
or intensity (data not shown), suggesting that the heat stability V72N and F74N in the core region of ␣A-crystallin had normal
of the protein was not affected by the mutation. activity (29). However, their conclusion was based on a single
It has been hypothesized that hydrophobic sites in ␣-crystal- assay conducted at 58 °C. We also found some activity of the
lin are responsible for chaperone-like activity (22, 23). How- mutant with ␤L-crystallin around this temperature. Also, un-
ever, this is not free of controversy (29, 47). In the present like Phe71, the Val72 and Phe74 residues show variations in
study we see a complete loss in the chaperone-like function of different vertebrate lens species. The conserved Phe71 residue
mutant ␣A-crystallin at and slightly above physiological tem- appears to be important for suppressing the aggregation of
peratures despite an increase in hydrophobicity. Smulders et proteins. Other factors like charge, hydrophobicity, and struc-
al. (29) observed an increase in the chaperone-like activity of an tural integrity may influence the functional property to differ-
␣AF74N mutant with a slight decrease in ANS binding. They ent extents. Recently, Kumar and Rao (53) produced a chimeric
concluded that there is no correlation between surface hydro- ␣-crystallin by swapping the domains of ␣A- and ␣B-crystallin
phobicity and chaperone-like activity. Experiments with super- and tested its effect on the chaperone-like activity. Interest-
␣A-crystallin have indicated that the disappearance of chaper- ingly, the ␣ANBC chimeric protein which contained residues
one-like activity may be independent of hydrophobicity (47). 1–79 of ␣A-crystallin, including a part of the functional site in
Further, Reddy et al. (18) have shown recently that hydropho- ␣A-crystallin, was completely inactive in suppressing the ag-
bicity is not the sole determinant of chaperone-like activity in gregation of insulin. However, the ␣BNAC chimeric protein
␣-crystallin. Recently, the studies with mini-␣A-crystallin containing the complete ADH binding sequence (25) and a part
showed that both hydrophobicity and ␤-sheet conformation of of the functional site in ␣A-crystallin had enhanced chaperone-
the functional element are essential for chaperone-like activity like activity. This suggests that other residues in the chaperone
(48). Although we see increased exposure of hydrophobic sur- site of ␣A-crystallin are also important in suppressing the
faces in the mutant, it is quite unlikely that all exposed hydro- aggregation of proteins. Therefore it would be interesting to
phobic patches would be involved in suppressing the substrate study the role of other conserved residues on chaperone-like
protein aggregation. We, as well as others, have observed bis- action.
ANS binding to residues other than those necessary for chap-
erone activity (27, 49). Taking these observations together, one Acknowledgments—We are grateful to Dr. P. D. Prasad for the help
can conclude that although hydrophobicity is important, the in site-directed mutagenesis and Jelena Kocergin in the expression and
purification of recombinant proteins.
extent of hydrophobicity does not reflect the chaperone-like
activity of the protein. REFERENCES
The ␣AF71G mutant has similar secondary structure to that 1. Harding, J. J., and Crabbe, M. J. (1984) in The Eye, pp. 207– 492, Academic
of the wild type. However, the tertiary structure shows some Press, New York
minor changes around the 285–295-nm region. The signal in 2. Groenen, P. J., Merck, K. B., de Jong, W. W., and Bloemendal, H. (1994) Eur.
J. Biochem. 225, 1–19
this region is produced by tyrosine or tryptophan residues. 3. Ingolia, T. D., and Craig, E. A. (1982) Proc. Natl. Acad. Sci. U. S. A. 79,
Because we did not observe any change in the tryptophan 525–529
F71G Mutant of ␣A-crystallin 47099
4. Merck, K. B., Groenen, P. J., Voorter, C. E., de Haard-Hoekman, W. A., Slingsby, C., Bloemendal, H., and De Jong, W. W. (1995) Eur. J. Biochem.
Horwitz, J., Bloemendal, H., and de Jong, W. W. (1993) J. Biol. Chem. 268, 232, 834 – 838
1046 –1052 30. Plater, M. L., Goode, D., and Crabbe, M. J. (1996) J. Biol. Chem. 271,
5. Bhat, S. P., and Nagineni, C. N. (1989) Biochem. Biophys. Res. Commun. 158, 28558 –28566
319 –325 31. Hepburne-Scott, H. W., and Crabbe, M. J. (1999) Mol. Vis. 5, 15–20
6. Dubin, R. A., Wawrousek, E. F., and Piatigorsky, J. (1989) Mol. Cell. Biol. 9, 32. Muchowski, P. J., Wu, G. J., Liang, J. J., Adman, E. T., and Clark, J. I. (1999)
1083–1091 J. Mol. Biol. 289, 397– 411
7. Nagineni, C. N., and Bhat, S. P. (1989) FEBS Lett. 249, 89 –94 33. Kumar, L. V., Ramakrishna, T., and Rao, C. M. (1999) J. Biol. Chem. 274,
8. Kato, K., Shinohara, H., Kurobe, N., Inaguma, Y., Shimizu, K., and Ohshima, 24137–24141
K. (1991) Biochim. Biophys. Acta 1074, 201–208 34. Horwitz, J., Bova, M., Huang, Q. L., Ding, L., Yaron, O., and Lowman, S.
9. van Noort, J. M., van Sechel, A. C., Bajramovic, J. J., el Ouagmiri, M., Polman, (1998) Int. J. Biol. Macromol. 22, 263–269
C. H., Lassmann, H., and Ravid, R. (1995) Nature 375, 798 – 801 35. Derham, B. K., van Boekel, M. A., Muchowski, P. J., Clark, J. I., Horwitz, J.,
10. Agius, M. A., Kirvan, C. A., Schafer, A. L., Gudipati, E., and Zhu, S. (1999) Acta Hepburne-Scott, H. W., de Jong, W. W., Crabbe, M. J., and Harding, J. J.
Neurol. Scand. 100, 139 –147 (2001) Eur. J. Biochem. 268, 713–721
11. Celet, B., Akman-Demir, G., Serdaroglu, P., Yentur, S. P., Tasci, B., van Noort, 36. Bova, M. P., Yaron, O., Huang, Q., Ding, L., Haley, D. A., Stewart, P. L., and
J. M., Eraksoy, M., and Saruhan-Direskeneli, G. (2000) J. Neurol. 247, Horwitz, J. (1999) Proc. Natl. Acad. Sci. U. S. A. 96, 6137– 6142
935–939 37. Shroff, N. P., Cherian-Shaw, M., Bera, S., and Abraham, E. C. (2000) Biochem-
12. Horwitz, J. (1992) Proc. Natl. Acad. Sci. U. S. A. 89, 10449 –10453 istry 39, 1420 –1426
13. Wang, K., and Spector, A. (1994) J. Biol. Chem. 269, 13601–13608 38. Cobb, B. A., and Petrash, J. M. (2000) Biochemistry 39, 15791–15798
14. Gopalakrishnan, S., Boyle, D., and Takemoto, L. (1994) Invest. Ophthalmol. 39. Horwitz, J., Huang, Q. L., Ding, L., and Bova, M. P. (1998) Methods Enzymol.
Vis. Sci. 35, 382–387 290, 365–383
15. Wang, K., and Spector, A. (1995) Invest. Ophthalmol. Vis. Sci. 36, 311–321 40. Augusteyn, R. C., Parkhill, E. M., and Stevens, A. (1992) Exp. Eye Res. 54,
16. van Boekel, M. A., de Lange, F., de Grip, W. J., and de Jong, W. W. (1999) 219 –228
Biochim. Biophys. Acta 1434, 114 –123 41. Siezen, R. J., Bindels, J. G., and Hoenders, H. J. (1980) Eur. J. Biochem. 111,
17. Datta, S. A., and Rao, C. M. (1999) J. Biol. Chem. 274, 34773–34778 435– 444
18. Reddy, G. B., Das, K. P., Petrash, J. M., and Surewicz, W. K. (2000) J. Biol. 42. Andrade, M. A., Chacon, P., Merelo, J. J., and Moran, F. (1993) Protein Eng. 6,
Chem. 275, 4565– 4570 383–390
19. Takemoto, L., and Boyle, D. (1994) Arch. Biochem. Biophys. 315, 133–136 43. Caspers, G. J., Leunissen, J. A., and de Jong, W. W. (1995) J. Mol. Evol. 40,
20. Rao, P. V., Huang, Q. L., Horwitz, J., and Zigler, J. S., Jr. (1995) Biochim. 238 –248
Biophys. Acta 1245, 439 – 447 44. de Jong, W. W., Caspers, G. J., and Leunissen, J. A. (1998) Int. J. Biol.
21. Carver, J. A., Nicholls, K. A., Aquilina, J. A., and Truscott, R. J. (1996) Exp. Macromol. 22, 151–162

Downloaded from www.jbc.org by guest, on October 28, 2010


Eye Res. 63, 639 – 647 45. de Jong, W. W., Leunissen, J. A., and Voorter, C. E. (1993) Mol. Biol. Evol. 10,
22. Raman, B., and Rao, C. M. (1994) J. Biol. Chem. 269, 27264 –27268 103–126
23. Das, K. P., and Surewicz, W. K. (1995) FEBS Lett. 369, 321–325 46. Derham, B. K., and Harding, J. J. (1999) Prog. Retin. Eye Res. 18, 463–509
24. Raman, B., and Rao, C. M. (1997) J. Biol. Chem. 272, 23559 –23564 47. van Rijk, A. F., van den Hurk, M. J., Renkema, W., Boelens, W. C., de Jong,
25. Sharma, K. K., Kaur, H., and Kester, K. (1997) Biochem. Biophys. Res. Com- W. W., and Bloemendal, H. (2000) FEBS Lett. 480, 79 – 83
mun. 239, 217–222 48. Bhattacharyya, J., and Sharma, K. K. (2001) J. Pept. Res. 57, 428 – 434
26. Sharma, K. K., Kaur, H., Kumar, G. S., and Kester, K. (1998) J. Biol. Chem. 49. Smulders, R. H., and de Jong, W. W. (1997) FEBS Lett. 409, 101–104
273, 8965– 8970 50. Lee, J. S., Satoh, T., Shinoda, H., Samejima, T., Wu, S. H., and Chiou, S. H.
27. Sharma, K. K., Kumar, G. S., Murphy, A. S., and Kester, K. (1998) J. Biol. (1997) Biochem. Biophys. Res. Commun. 237, 277–282
Chem. 273, 15474 –15478 51. Santhoshkumar, P., and Sharma, K. K. (2001) Mol. Vis. 7, 172–177
28. Sharma, K. K., Kumar, R. S., Kumar, G. S., and Quinn, P. T. (2000) J. Biol. 52. Srinivas, V., Datta, S. A., Ramakrishna, T., and Rao, C. M. (2001) Mol. Vis. 7,
Chem. 275, 3767–3771 114 –119
29. Smulders, R. H., Merck, K. B., Aendekerk, J., Horwitz, J., Takemoto, L., 53. Kumar, L. V., and Rao, C. M. (2000) J. Biol. Chem. 275, 22009 –22013

Das könnte Ihnen auch gefallen