SlideShare a Scribd company logo
1 of 15
Download to read offline
proteinsSTRUCTURE O FUNCTION O BIOINFORMATICS
Analysis of the structure of human
apo-S100B at low temperature indicates a
unimodal conformational distribution is
adopted by calcium-free S100 proteins
Shahid Malik,1
Matthew Revington,1
Steven P. Smith,2
and Gary S. Shaw1
*
1 Department of Biochemistry, The University of Western Ontario, London, Ontario, N6A 5C1, Canada
2 Department of Biochemistry, Queen’s University, Kingston, Ontario, K7L 3N6, Canada
INTRODUCTION
The S100 protein family makes up one of the largest EF-hand
groups of proteins, comprising at least 25 members in
humans.1,2 These low molecular weight (21–24 kDa) proteins
are generally expressed in a cell-specific manner having a variety
of biological roles including the regulation of cytoskeletal protein
assembly,3–7 control of protein phosphorylation,8–13 and modu-
lation of the activities of enzymes such as aldolase14,15 and
phosphoglucomutase.16 To date over 90 potential protein targets
have been identified for the S100 proteins,1 which include the
annexins (A1, A2, A5, A6, A11),17–21 the ubiquitination protein
CacyBP,22,23 F-actin,24 p53,25–27 and tau.28 Most of these
interactions are controlled through calcium binding and a subse-
quent conformational change of the S100 protein, revealing a
hydrophobic surface that recognizes the target protein sequence.
In general, this hydrophobic binding surface can also be probed
through binding of the calcium-bound S100 protein to phenyl
sepharose or anilinonaphthalene-8-sulfonic acid (ANS). These
probes have proven to be a connecting trademark for the S100
proteins, placing them within the calcium ‘‘sensor’’ group of EF-
hand proteins that also includes the muscle protein troponin-C
and the multifunctional protein calmodulin.29,30 Further interest
in the S100 members has evolved because some of these proteins,
such as S100B and S100A8/S100A9 show elevated expression lev-
els in Alzheimer’s disease31,32 and rheumatoid arthritis,33,34
respectively, suggesting they may have roles in these diseases.
S100B is one of the best-characterized members of the S100
protein family. As with all other S100 proteins, except S100G
(calbindin D9k), S100B can form homodimers, comprising two
noncovalently associated 91-residue subunits. Further, yeast two-
hybrid and optical biosensor experiments have indicated that
Abbreviations: nOe, nuclear overhauser effect; RDC, residual dipolar couplings; rmsd, root
mean square deviation.
Grant sponsor: Canadian Institutes of Health Research (GSS).
*Correspondence to: Gary S. Shaw, Department of Biochemistry, The University of Western
Ontario, London, Ontario, N6A 5C1, Canada. E-mail: gshaw1@uwo.ca.
Received 30 November 2007; Revised 8 February 2008; Accepted 19 February 2008
Published online 2 April 2008 in Wiley InterScience (www.interscience.wiley.com).
DOI: 10.1002/prot.22037
ABSTRACT
S100B is one of the best-characterized members of the
calcium-signaling S100 protein family. Most S100
proteins are dimeric, with each monomer containing
two EF-hand calcium-binding sites (EF1, EF2). S100B
and other S100 proteins respond to calcium increases
in the cell by coordinating calcium and undergoing a
conformational change that allows them to interact
with a variety of cellular targets. Although several
three dimensional structures of S100 proteins are
available in the calcium-free (apo-) state it has been
observed that these structures appear to adopt a wide
range of conformations in the EF2 site with respect to
the positioning of helix III, the helix that undergoes
the most dramatic calcium-induced conformational
change. In this work, we have determined the struc-
ture of human apo-S100B at 108C to examine
whether temperature might be responsible for these
structural differences. Further, we have used this
data, and other available apo-S100 structures, to
show that despite the range of interhelical angles
adopted in the apo-S100 structures, normal Gaussian
distributions about the mean angles found in the
structure of human apo-S100B are observed. This
finding, only obvious from the analysis of all avail-
able apo-S100 proteins, provides direct structural evi-
dence that helix III is a loosely packed helix. This is
likely a necessary functional property of the S100 pro-
teins that facilitates the calcium-induced conforma-
tional change of helix III. In contrast, the calcium-
bound structures of the S100 proteins show signifi-
cantly smaller variability in the interhelical angles.
This shows that calcium binding to the S100 proteins
causes not only a conformational change but results
in a tighter distribution of helices within the EF2 cal-
cium binding site required for target protein interac-
tions.
Proteins 2008; 73:28–42.
VVC 2008 Wiley-Liss, Inc.
Key words: calcium-binding proteins; structure; nmr.
28 PROTEINS VVC 2008 WILEY-LISS, INC.
S100B can form heterodimers in the cell with S100A1,
S100A6, and S100A11.35,36 Each S100B monomer com-
prises a globular domain containing two EF-hand cal-
cium-binding motifs. Sequence comparison of the S100
proteins reveals the presence of a basic N-terminal
pseudo (C) calcium-binding site, comprising 14 residues
(site I) between helices I and II (EF1), and an acidic 12-
residue canonical calcium-binding site in the C-terminus
(site II) flanked by helices III and IV (EF2). The region
joining the two calcium-binding sites (linker), and the
extreme N- and C-terminal regions of the S100 family
members are more divergent in their sequences.37 Three-
dimensional structures of S100A1038 and S100A1139
show that residues within these latter three regions are
important for the interactions with the annexin proteins
that are thought to be important in membrane vesicula-
tion processes.
Three-dimensional structures of several S100 proteins
in the calcium-free (apo) and calcium-bound states,
including S100B,40–43 S100A644–46 and S100A11,39,47
shows that they undergo a significant conformational
change on binding calcium, exposing several hydrophobic
residues that are required for interacting with specific bi-
ological target molecules. For example, the conforma-
tional change in S100A11 indicates a 408 reorientation of
helix III occurs on calcium binding leading to a more
open or exposed conformation. In other S100 proteins,
such as S100B and S100A6, the calcium-induced confor-
mational change ranges between $608 and $908. These
observations might indicate that the calcium-induced
structural change is different for each S100 protein. Alter-
natively, the structures could reflect a wide range of con-
formational space that can be sampled by helix III in ei-
ther the apo- or calcium-bound structures. For example,
the unstable nature of helix III in apo-S100B is consistent
with amide protons in this helix that exchange up to
three orders of magnitude faster than those in helices I,
II, or IV.48 In principle, a distinction between different
orientations for helix III (and therefore different confor-
mational changes) or multiple orientations that sample a
broad range of conformations should be possible through
a detailed analysis of existing or new S100 structures. To
date, however, this has not been done.
In this work, we have determined the water-refined,
three-dimensional structure of the calcium-free human
S100B dimer at low temperature (108C) to define
whether major temperature-induced structural changes
occur compared to other S100 structures determined at
higher temperatures. We have also used chemical shift in-
formation and residual dipolar coupling experiments to
determine if more subtle temperature-induced structural
changes occur in apo-S100B, and in particular helix III.
Using this data, and the structure of human apo-S100B,
we have assessed the interhelical interactions and super-
positions for the 11 calcium-free S100 protein structures
currently available. We find that the apo-S100 proteins
exhibit normal Gaussian distributions expected for a sin-
gle conformational distribution. These results provide
evidence that the structures of apo-S100 proteins reflect
their ability to survey a wide range of conformational
space rather than adopting a specific conformation. The
structure of human apo-S100B is close to the prototypi-
cal structure in this population.
MATERIAL AND METHODS
Sample preparation
Uniformly 15
N- and 13
C-labeled recombinant human
S100B was expressed in Escherichia coli strain N99, and
purified to homogeneity as previously described.49 All
NMR experiments utilized 1–2 mM human S100B sam-
ples dissolved in 90% H2O/10% D2O containing 50 mM
KCl at pH 7.19. DSS was used as an internal standard. A
single 15
N/13
C-labeled sample was used for all triple reso-
nance NMR experiments.
NMR spectroscopy
All NMR spectroscopy utilized Varian 600 and 800
MHz spectrometers with pulse field gradient triple reso-
nance probes. Data for the backbone sequential assign-
ments was collected at 108C and included HNCA,50
HNCACB,51 CBCA(CO)NH,52 HNCO, and 1
H-15
N
HSQC experiments. Side chain assignments were made
from C(CO)NH, HC(CO)NH,53 HCCH-COSY54 (90%
D2O/10% H2O), HBCBCGCDHD,55 and 1
H-13
C HSQC
experiments. Interproton distances were measured from
15
N-NOESY (mixing time 150 ms), 13
C-NOESY (mixing
time 100 ms, in 90% D2O/10% H2O) and F1-filtered, F3-
edited NOESY56 experiments (mixing time 100 ms) were
collected at 800 MHz at the Canadian National High
Field NMR Centre (NANUC). Slow-exchanging amide
protons were identified via a series of 1
H-15
N HSQC
experiments taken at time intervals (5 min, 5 h, 10 h,
and 16 h) after dissolving the protein in D2O. 1
H-15
N re-
sidual dipolar couplings were measured from a series of
IPAP-HSQC experiments57–59 using a 0.7 mM sample
of 15
N-labeled apo-S100B in 12 mg/mL Pf1 phage with
50 mM KCl at pH 7.19. Spectra were acquired at 35, 25,
15, 10, and 58C. An identical sample lacking phage was
used to measure isotropic 1
H-15
N splittings at the same
temperatures. All NMR spectra were processed using
NMRPipe60 and analyzed using NMRView61 software.
Structure calculations
Three-dimensional structures of apo-S100B were calcu-
lated using the simulated annealing protocol in the pro-
gram CNS62 with the use of noncrystallographic symme-
try. A total of 2504 nOes, including 1328 intraresidue,
556 sequential, 370 short-range, 160 long-range, and 90
intermolecular distances were used. After the initial
Conformational Distribution for Apo-S100 Proteins
PROTEINS 29
structure calculation was completed to define the second-
ary structure of the protein, 124 hydrogen bond distance
restraints and 336 dihedral restraints were included in
the calculations. Interproton distances for all the proton
pairs, except intermolecular nOes, were calibrated using
maximum and minimum nOe intensities for known
DNHa distances.63 For all intermolecular nOes, an upper
distance of 5 A˚ was used. Dihedral angle restraints were
determined using the TALOS protocol64 based on chemi-
cal shift data from Ha, Ca, Cb, and CO resonances of
apo-S100B. Angular restraints were only selected when 9/
10 dihedral matches fell within an allowed region of the
Ramachandran plot. The resulting w and / angles were
restricted to two times the error from the TALOS output
for the structure calculation with a minimum error of
Æ208. Hydrogen bonds were identified from slowly
exchanging NH resonances at the 16 h time point from
the incubation of apo-S100B in D2O and where tempera-
ture coefficients > 24.5 ppb/K could be measured. For
each hydrogen bond distance restraints for NHÀÀO (1.8–
2.3 A˚ ) and NÀÀO (2.3–3.3 A˚ ) were used. The final water
refinement utilized the RECOORD method65 for CNS.
Structures were viewed and analyzed using MOLMOL.66
VADAR67 (University of Alberta) was used to calculate
the accessible surface area for all proteins described. All
interhelical angles were calculated using the Vector Ge-
ometry Mapping method.29,68 In cases where NMR
structures were used, calculations were done for multiple
structures and the average Æ standard deviation reported.
Pairwise comparisons and rmsd calculations for all apo-
S100 structures were done using the program Chimera.69
The atomic coordinates for the 20 lowest energy struc-
tures of human apo-S100B have been deposited in the
RCSB under accession number 2PRU.
RESULTS AND DISCUSSION
The backbone and side chain resonance assignments
for apo-S100B at 108C and 358C were determined by the
acquisition and analysis of standard two- and three-
dimensional NMR experiments. In total, assignments of
1
H, 13
C, and 15
N resonances for 90 of the 91 residues in
apo-S100B were completed. The aromatic side-chain res-
onances for all seven phenylalanine, one tyrosine, and
four histidine residues in each apo-S100B subunit were
accomplished and were critically important for the struc-
ture determination of the protein since a large number
of intra- and intermolecular nOe correlations were
observed to these residues.
Temperature dependence of apo-S100B
The majority of three-dimensional structures for apo-
S100 proteins have been completed using NMR spectros-
copy at temperatures above 258C. For example, data used
for the structures of apo-S100A6,37,44 S100B40 and
S100P70 was acquired at 25–388C. In contrast, the only
two X-ray structures of calcium-free S100 proteins, apo-
S100A371 and apo-S100A646 had data collected at
21738C, although crystals were grown near ambient
temperatures. To determine the influence of temperature
on the apo-S100B structure a series of 1
H-15
N HSQC
spectra were collected at temperatures ranging between
10–358C. As shown in Figure 1 the majority of the amide
correlations shift downfield in both 1
H and 15
N dimen-
sions as the temperature is lowered. In general, most of
the resonances exhibited some line broadening at the
lower temperatures, which is consistent with a longer
correlation time due to an increased viscosity/tempera-
ture ratio. Further, analysis of the temperature-dependent
changes in the cross peak position showed that each res-
onance shifted in a linear fashion. The largest change in
amide proton chemical shift was $0.2 ppm as the tem-
perature was lowered. Similar temperature dependent
changes in 1
H chemical shift have been noted for the EF-
hand protein troponin-C72 where a large structural
change has not been noted. In addition, the changes in
chemical shift with decreased temperature for apo-S100B
are much less than the 1
H chemical shift changes on cal-
cium-binding or interaction with other proteins73,74
where a large structural reorganization of the protein
occurs. The temperature coefficients for each amide reso-
nance (2DdHN/T) are shown in Figure 2. The data shows
Figure 1
1
H-15
N HSQC spectra for human apo-S100B shown as a function of
temperature. The spectra are superimposed and plotted using filled contours for
108C and a single contour levels at 25 and 358C. Assignments for most of the
isolated correlations are indicated near the 108C spectrum. Arrows are used to
show the direction of movement for some of the most affected resonances over
the 358C to 108C temperature range.
S. Malik et al.
30 PROTEINS
the average temperature coefficient is 23.2 ppb/K and
most of the residues in apo-S100B are found within one
standard deviation (Æ2.3 ppb/K) of this value. Since
there are no regions of the sequence that exhibit anoma-
lous coefficients this further indicates that large tempera-
ture-induced structural changes in apo-S100B do not
occur as the temperature is lowered. It has been shown
for bovine pancreatic trypsin inhibitor and lysozyme that
temperature coefficients more positive than 24.5 ppb/K
indicate an amide proton that is hydrogen bonded.75
This appears to hold true for many residues within the
helical regions of apo-S100B. However, several amide
protons lie in the lower portion of [Fig 2(A)], below the
24.5 ppb/K threshold, indicating poorer hydrogen bond-
ing. Many of these are located in unstructured regions,
including those that lie either at the extreme C-terminus
of the protein (E91) or within calcium-binding sites I
(G19, G22, D23, K24, S30) and II (D61, G64, E67, F70).
It is interesting to note that two amide resonances that
shift greater than one standard deviation from the mean
lie at the observed S100 dimerization interface involving
helix I (L3, L10). While a loss of intramolecular hydrogen
bonding could cause this effect, this result may arise
from a minor temperature dependent alteration of the
helix I-helix I0
interface. An additional significant change
was noted at the N-terminus of helix III (I47, Q50), a
region that contains some of the fastest exchanging am-
ide protons in apo-S100B.48
To assess whether any re-orientation of helices in apo-
S100B occurred between 358C and 108C, we measured
HN residual dipolar couplings using two-dimensional
IPAP 1
H-15
N HSQC experiments at both temperatures
[Fig. 2(B)]. Residual dipolar couplings provide orienta-
tional information about the NÀÀH bond vector in a
partially oriented sample in a magnetic field. The figure
shows the expected clustering of negative couplings for
helices I and IV since these helices are nearly coplanar as
shown in previously determined S100 structures.40 Over-
all, at the two temperatures studied most residual dipolar
couplings changed by less than 1 Hz, which is close to
the measurement error for the experiments. This is in
contrast to much larger changes observed for the ribonu-
clease S-peptide, where a significant temperature-induced
structural alteration occurs. In apo-S100B, the residual
dipolar coupling experiments suggest that little reorienta-
tion of any of the helices occurred over the 10–358C
temperature range. Alternatively, it is possible that these
experiments report only the average position of the heli-
ces in apo-S100B and would not be sensitive to helix
movement or helices that occupy a range of conforma-
tions. To determine this latter case, a more detailed ex-
amination of the apo-S100B structure at lower tempera-
ture and its comparison with other apo-S100 structures
was completed.
Description of the structure of
human apo-S100B
The three-dimensional structure of human apo-S100B
was determined at 108C using a total of 2504 nOes, 336
dihedral restraints, and 124 hydrogen bonds. Hydrogen
bonds and dihedral restraints were only used in a-helical
regions clearly defined from nOe patterns. The amide
protons involved in hydrogen bonding were identified
from their slow exchange with D2O and measured tem-
perature coefficients > 24.5 ppb/K. The 20 lowest energy
structures obtained from CNS calculations were used for
further refinement using a full nonbonded potential to-
gether with explicit solvent molecules. While use of this
approach made little difference to the overall structure,
Figure 2
A: 1
H temperature coefficients for the backbone amide (HN) resonances of
human apo-S100B. The coefficients were determined from the slope of a plot of
HN chemical shift vs. temperature between 308 K and 283 K and plotted. The
average 1
H temperature coefficient was 23.2 ppb/K (dashed line). Temperature
coefficients larger than 24.5 ppb/K (solid line) are consistent with hydrogen
bonding. Residues that exhibited temperature coefficients more negative than
24.5 ppb/K, are not expected to be hydrogen bonded and are labeled for clarity.
B: HN residual dipolar coupling constants for human apo-S100B measured at
358C (*) and 108C (n). Only residues that exhibited clearly resolved HN
correlations were measured. Bars above each figure are shown to represent the
positions of the four a-helices, determined in the current work.
Conformational Distribution for Apo-S100 Proteins
PROTEINS 31
improvements were noted in the average nOe and angu-
lar restraint violations, and their corresponding energies.
In addition, the number of close contacts dropped by
nearly 70% in the water-refined approach, a similar result
to that noted for other systems.65 The ensemble of the
20 lowest-energy structures of human apo-S100B is
shown in Figure 3 and structural statistics are listed in
Table I. Structures were chosen based on their low ener-
gies and contained no distance violations greater than
0.5 A˚ and no angle violations greater than 58. The family
of structures shows two S100B monomers, each with
four well-defined helices (Fig. 3: I–IV, I0
–IV0
). The sym-
metric relationship between the monomers is evident
from a twofold rotational axis passing through the dimer
interface approximately perpendicular to helices I and I0
,
and parallel to helices IV and IV0
. The NMR spectra were
consistent with this symmetric nature as one resonance
was observed for most peaks. Some peaks, such as L3
exhibited multiple peaks (see Fig. 1) because of the pres-
ence of formyl- and desformyl methionine forms of the
protein that have been previously characterized.76 In all
cases, these peaks shifted nearly identically during tem-
perature studies (see Fig. 1) indicating the differential N-
terminal modification has little affect on dimerization.
The rmsd of all helices in the dimer relative to the mean
structure was 0.55 Æ 0.07 A˚ for the backbone and 1.03
Æ 0.09 A˚ for all heavy atoms.
Each S100B monomer is formed by two helix-loop-he-
lix EF-hand calcium-binding motifs joined by a linker
region [Fig. 3(A)]. The N-terminal EF-hand consists of
helix I (E2-Y17), calcium binding loop I (S18-K28), and
helix II (K29-E39). The C-terminal EF-hand comprises
helix III (Q50-D60), calcium-binding loop II (D61-D69),
and helix IV (F70-T81). Each of the four helices is well
defined in all 20 low-energy structures, with the rmsd
values 0.33 Æ 0.08 A˚ for helix I, 0.23 Æ 0.07 A˚ for helix
Figure 3
Three-dimensional structure of human apo-S100B determined at 108C. A: Ribbon diagram of apo-S100B showing helices I and I0
(blue), II and II0
(yellow), III and III0
(pink), and IV and IV0
(red). B: Superposition (N, Ca, C0
) of the family of 20 low-energy structures of apo-S100B obtained after water-refinement shown in the same
orientation as (A).
Table I
Structural Statistics of Apo-S100B
Restraints for RECOORD refined structure calculation
Total NOEs 2504
Intraresidual NOEs 1328
Sequential NOEs 556
Long-range NOEs 530
Intermolecular 90
Dihedral angles /; w (TALOS) 336
Hydrogen bonds 124
Energies (kcal/mol)
ETotal 28066.5 Æ 224.8
ENOE 0.81 Æ 0.10
Echid 0.83 Æ 0.50
EL-J 2739.6 Æ 35.2
Structure quality
r.m.s.d. from experimental restraints
Angular (8) 0.1798 Æ 0.0357
Distance (Š) 0.0173 Æ 0.0015
r.m.s.d. from idealized geometry
Bond (Š) 0.0051 Æ 0.0004
Bond angles (8) 1.3380 Æ 0.0637
Improper torsion (8) 1.7632 Æ 0.1050
Ramachandran plot statistics
Residues in favored regionsa
94.9%
r.m.s.d. to mean structure (Š)b
Backbone atoms 0.55 Æ 0.07
Heavy atoms 1.03 Æ 0.09
Bad contacts (avg. per structure) 2.4
a
Reflects residues in both most favored and additionally favored regions.
b
Precision is calculated for residues found in helices I (3–16), II (29–39), III
(50–60), and IV (70–80) in apo-S100B.
S. Malik et al.
32 PROTEINS
II, 0.14 Æ 0.05 A˚ for helix III, and 0.17 Æ 0.03 A˚ for he-
lix IV. For the calcium-binding loop regions, long-range
nOes were found for residues L27 with C68 and D69, K28
with E67 and C68, and K29 with C68 indicating the two
calcium-binding loops were in close proximity. Evidence
from nOe data also showed that extensive hydrophobic
interactions occurred between the four helices. These
interactions include contacts between V13 and F14 (helix
I) with L35 (helix II), S30, L32, and I36 (helix II) with
V56, M57, and L60 (helix III), V52, V56 (helix III) with
F76, M79 and V80 (helix IV), L10 (helix I) with residue
F73 (helix IV), and L32, L35, I36 (helix II) with F73 and
V80 (helix IV). Analysis of the family of structures indi-
cated that 94.9% of all the residues were in the allowed
regions of the Ramachandran plot.
The dimerization of apo-S100B occurs through an anti-
parallel alignment of helices I-I0
and IV-IV0
, and the per-
pendicular association of these pairs of helices to form an
X-type bundle. Residues that define the dimer interface in
human apo-S100B were identified from a 13
C F1-filtered,
F3-edited NOESY spectrum using 13
C, 15
N-labelled and
unlabelled proteins in a 50:50 ratio in solution. Unambig-
uous intermolecular nOes from this data showed interac-
tions between A6 (helix I) and A60
, A90
and L100
(helix I0
),
and L3 (helix I) with V130
(helix I0
). The perpendicular
association of helices I and IV0
was evident from unambig-
uous intermolecular nOes between L3 and M7 (helix I)
with V770
and T810
(helix IV0
), respectively. The orienta-
tion of helices IV and IV0
was determined by nOes between
M74 (helix IV) with M740
(helix IV0
) and, F70 (helix IV)
with T820
and F870
(helix IV0
). Further hydrophobic inter-
actions at the dimer interface were observed between the
N-terminus of helix I (L3), and the N-terminus of the
linker region in the partner monomer (L400
).
The three-dimensional structures of rat40 and bovine43
apo-S100B have been previously determined using NMR
spectroscopy. A comparison of these structures reveals a
backbone rmsd of 2.00 A˚ (residues 2–84) between the two
structures indicating there are some significant differences
between these two proteins. This arises mostly from dissimi-
larities in the orientation of helix III and calcium-binding
loop II. While it is possible sequence variation could con-
tribute to this difference, this seems unlikely given that rat
and bovine sequences share 95.6% identity (87 of 91 resi-
dues). Further, these proteins have identical helix III
sequences and a single minor change in calcium-binding
loop II (E62 in rat; S62 in bovine). The current structure of
human apo-S100B, which shares 97.8% and 96.7% identi-
ties with rat and bovine proteins, respectively, allows some
of the differences between the rat and bovine structures to
be resolved. Using the helices of these apo-S100 proteins for
comparison, there is a backbone rmsd of 1.76 A˚ between
human and rat apo-S100B that increases to 2.29 A˚ between
the human and bovine proteins. These inequities largely
result from differences in the positions of helices III and IV
where an rmsd of 1.24 A˚ exists between human and rat pro-
teins but increases to 2.32 A˚ between human and bovine
proteins. Only a single minor, conservative residue change
occurs between human and bovine proteins at position 80
(I80 in bovine; V80 in human) in helix IV indicating the
sequence difference probably does not account for the larger
variance in helix positions between these proteins. Thus,
when comparing the a-helices it is clear that human apo-
S100B is more similar to the rat S100B structure than the
bovine structure. The similarity between the orientation of
a-helices in human and rat apo-S100B proteins becomes
more apparent when interhelical angles are considered
(described later).
The human apo-S100B structure (see Fig. 3) indicates
that calcium-binding loops I (G19-E32) and II (D61-
E72) are less well defined (backbone rmsd 2.27 and
1.18 A˚ , respectively) than the helices. A similar observa-
tion has been made for bovine apo-S100B43 although
calcium-binding loop I is the better defined of the two
loops in that protein. However, in rat apo-S100B,40 the
backbone rmsds for calcium-binding loops I (0.22 A˚ )
and II (0.15 A˚ ) are similar to those found in the a-heli-
ces suggesting the loops adopt a tight structure with lim-
ited flexibility. In human apo-S100B, a network of nOes
similar to those observed for helices I-IV was not evident
in calcium-binding loops I and II, contributing to its
poorer definition. Further, amide exchange experiments
show that most of the amide protons within calcium-
binding loops I and II have protection factors up to six
orders of magnitude lower than amide protons in the a-
helices.48 This indicates that a poorer series of hydrogen
bonds exists in the calcium-binding loops of human apo-
S100B and that most of the amide protons are exposed
to solvent. In addition, 15
N relaxation experiments77
have shown that the order parameters (S2
) for residues
in calcium-binding loops I and II average 0.80, perhaps
indicating a greater degree of flexibility within the loops
than the helices (S2
5 0.84–0.87). It is interesting that
15
N relaxation experiments with another EF-hand cal-
cium-binding protein troponin-C,78,79 have indicated
calcium-binding sites I and II in the N-terminal domain
have lower order parameters also, characteristic of greater
flexibility. Thus, the poor definition of the calcium-bind-
ing loops in human apo-S100B is most consistent with
flexibility of the backbone within these regions. This
interpretation is in agreement with crystallographic stud-
ies of apo-S100A371 and apo-S100A646 where thermal
factors about two-times those found in the helices have
been observed for site I in apo-S100A6 and site II in
apo-S100A3, corresponding to backbone atomic displace-
ments near 0.6 A˚ .
The apo-S100 proteins show variations in
interhelical angles
All S100 proteins, with the exception of S100A10,
undergo a calcium-induced conformational change that
Conformational Distribution for Apo-S100 Proteins
PROTEINS 33
allow for their interaction with a variety of target pro-
teins. In general, this change involves a reorientation of
helix III and repacking of helix II that results in the ex-
posure of a broad, hydrophobic surface. To assess the
details of this conformational change and to determine
whether specific S100 proteins adapt to calcium binding
in different manners, knowledge of the structures of
many S100 protein family members is required in the
apo- and calcium-bound states. We have used the struc-
ture of human apo-S100B to assess the detailed structure
of all S100 proteins in the calcium-free state to under-
stand the first part of the calcium-binding response.
The structure of human apo-S100B determined at
108C was compared with NMR and X-ray crystallo-
graphic structures of S100A1, S100A3, S100A4, S100A6,
S100A11, S100A13, bovine and rat S100B, and S100P
using a variety of criteria, including interhelical distances
and angles, helix rotation and buried surface area. The
helix–helix relationship in this broad range of S100 struc-
tures was examined using the vector geometry mapping
(VGM) method, which provides important information
about the tip angle (y) between helices, the helix projec-
tion angle (/) and the helix role (x), and gives the best
overall picture of the spatial arrangement of helices.29,68
In our assessment, first shown for apo-calmodulin, we
chose helices adjacent to the calcium-binding loops such
that the incoming helix ended at the hydrophobic residue
immediately preceding the calcium-binding loop. In apo-
S100B this corresponded to Y17 and L60 in helices I and
III. The exiting helix started two residues before the
bidentate glutamate-coordinating residue that terminates
the calcium-binding loop (E31 and E72 in helices II and
IV of human S100B). In all cases, a careful examination
was made of the helix selections by shifting the helix des-
ignation by one or two residues to insure the y and /
were not grossly affected. It was observed that some heli-
ces, especially helices I and III, were very sensitive to this
selection due to the presence of some helix twist or
bend.
The interhelical angles y and / for human apo-S100B
and 10 other apo-S100 structures are listed in Table II
and shown graphically in Figure 4. The y and / angles
between helices I and II for the pseudo EF-hand (EF1) in
human apo-S100B are 72 Æ 48 and 98 Æ 88, respectively.
These are in excellent agreement with the average angles
(63 Æ 108, 99 Æ 78) obtained when all other apo-S100
structures are considered, indicating that helices I and II
adopt near identical orientations amongst all available
apo-S100 protein structures. This is shown clearly in
[Fig. 4(A)], where there is a clustering of the exiting he-
lix II for each structure. In addition, the pairwise super-
positions for helices I and II in these structures (55 com-
binations) was binned and fit with a Gaussian distribu-
tion having a midpoint of 1.46 Æ 0.56 A˚ [Fig. 5(A)].
This data passed multiple tests for a normal distribution.
The excellent agreement of the midpoint for the Gaussian
fit, the median (1.46) and positive normal distribution
tests was strong evidence for a single structural popula-
tion. This would indicate that the distribution of angles
for helices I and II in the apo-S100 structures is most
consistent with a single conformational family. Further,
the range of angles for EF1 in the apo-S100 proteins (Ta-
ble II) is considerably tighter than that observed for heli-
ces I and II in all EF-hand protein structures (1158).80
A comparison of helices I and II in EF1 from the S100
family to apo-calmodulin shows that calmodulin has a
significantly smaller tip angle (y) of 46 Æ 28, and larger
horizontal angle (/) of 123 Æ 38, both of which lie out-
side the ranges observed for any single S100 protein. As
shown in Figure 4, the position of the C-terminus of he-
lix II is closer to the N-terminus of helix I for apo-cal-
modulin than for any of the S100 proteins, resulting in y
and / angles for apo-calmodulin that differ by about
2178 and 248, respectively, when compared with the av-
erage apo-S100 structure. Human apo-S100B and the rest
of the S100 proteins thus have a more open conforma-
tion for EF1 than apo-calmodulin, a result reaffirmed
when considering interhelical distances between helices I
and II of EF1. The separation of these helices in the S100
proteins is approximately 3.6 A˚ greater than those in
apo-calmodulin.
The tip and horizontal angles for EF2 of human apo-
S100B are y 5 25 Æ 28 and / 5 2146 Æ 88, respectively
(Table II). These are well within the range observed for
all apo-S100 proteins (y 5 25 Æ 88, / 5 2142 Æ 468).
In general the exiting helix IV in the apo-S100B proteins
is angled in an opposite direction compared with apo-
calmodulin (see Fig. 4), which has a positive / angle
most similar to S100P. Unlike EF1, the larger deviations
in the y and / angles for EF2 indicate that potential dif-
ferences might exist between the structures due to
sequence differences, helix–helix interactions or both.
Alternatively, the differences in the structures could
reflect a lack of definition in the NMR structures due to
limited numbers of nOes between helices III and IV. This
does not appear to be the case since most of the apo-
S100 structures have tight ranges of y and / angles
( 208) indicating the nOe distance information used in
these structures was sufficient. In addition, variation is
noted for the two x-ray structures that are available
(S100A3 and S100A6) that have / angles that differ by
>208, although this difference is clearly smaller than that
exhibited by the apo-S100 structures determined by
NMR spectroscopy.
An examination of the y and / angles for the apo-
S100 structures shows that, with the exception of S100P,
all the apo-S100 structures are found within 88 of the av-
erage y angle with approximately equal numbers on ei-
ther side of the average. This indicates that the opening
angles between helices III and IV are similar as noted by
the mostly parallel arrangement of the helices for the
apo-S100 proteins [Fig. 4(D)]. On the other hand, the
S. Malik et al.
34 PROTEINS
TableII
EF-HandAnglesofApo-S100Proteinsa
Protein
EF1b
EF2
N-terminalcoordinate
ofsecondhelix
y
(degree)
/
(degree)
x
(degree)
N-terminalcoordinateof
secondhelix
y
(degree)
/
(degree)
x
(degree)
hS100Bc
11.8Æ0.7,23.3Æ1.7,21.1Æ0.772Æ498Æ8129Æ88.4Æ1.1,11.5Æ0.7,21.8Æ0.525Æ22146Æ883Æ8
bS100Bd
10.3Æ1.7,1.6Æ1.5,25.4Æ2.153Æ2105Æ493Æ41.4Æ1.3,10.8Æ1.4,28.2Æ1.310Æ42124Æ1045Æ3
rS100Be
12.3Æ0.3,24.0Æ1.2,25.2Æ0.758Æ193Æ3129Æ47.8Æ0.6,10.4Æ0.5,24.6Æ0.515Æ12127Æ595Æ4
S100A1f
13.1Æ1.3,23.3Æ2.4,24.7Æ1.368Æ391Æ5166Æ63.3Æ1.9,14.3Æ0.5,20.3Æ0.333Æ1284Æ473Æ8
S100A3g
11.2,20.6,24.25510212911.7,10.9,23.0262116107
S100A4h
12.4Æ1.2,22.2Æ1.4,23.2Æ0.768Æ2105Æ3114Æ48.1Æ2.8,11.9Æ2.1,23.7Æ1.419Æ22151Æ770Æ10
#
S100A6i
10.1Æ0.7,0.4Æ0.3,24.8Æ0.556Æ1103Æ3113Æ310.1Æ0.8,11.0Æ0.3,22.0Æ0.434Æ22161Æ395Æ3
S100A6j
10.361,20.181,24.8095610512111.4,11.4,23.5242138111
S100A11k
9.9Æ0.4,21.3Æ0.7,24.3Æ0.363Æ2101Æ1123Æ88.7Æ2.9,6.9Æ1.3,22.2Æ1.230Æ4163Æ1073Æ20
S100Pl
10.8Æ0.7,21.0Æ0.7,25.9Æ1.154Æ3105Æ4104Æ29.0Æ0.2,1.49Æ0.7-4.7Æ0.570Æ3121Æ386Æ4
S10013m
13.5Æ1.5,28.3Æ2.5,20.3Æ1.284Æ684Æ9180Æ92.2Æ1.4,14.3Æ0.5,20.5Æ1.030Æ3279Æ757Æ6
AverageS10063Æ1099Æ7128Æ2529Æ15(25Æ8)2142Æ4681Æ20
CaM10.6Æ0.6,20.7Æ0.7,23.2Æ0.346Æ2123Æ3104Æ510.9Æ0.9,5.6Æ1.6,21.8Æ1.048Æ4139Æ489Æ4
a
AngleswerecalculatedusingVectorGeometryMapping(VGM)methodusingapo-calmodulin(PDBentry1DMO)asthereference.29,67
b
HelicesdefinedforEF1arehelixIandIIandforEF2ashelixIIIandIVfor:S100B-L10–Y17,K29–E39,V53–L60,F70–V80;S100A1-L11–H18,K30–E40V54–L61,F71–L81;S100A3-I12–Y19,Q31–E41Y55–L62,F72–L82;
S100A4-M12–Y19,K31–E41F55–L62,F72–I82;S100A6-L12–Y19,K31–L41,I53-L60(#
E52–D59),F70–L80;S100A11-L13–Y20,K32–E42,L56–L63,F73–L83;S100P-I11–Y18,K30–E40V54–L61,F71–I81;S100A13-V16-F23,
V35-Q45,L56-L63,F73-L83.
PDBentriesare
c
2PRU
d
1CFP
e
1B4C
f
1K2H
g
1KSO
h
1M31
i
2CNP
j
1K9P
k
1NSH
l
1OZO
m
YUS.
Conformational Distribution for Apo-S100 Proteins
PROTEINS 35
differences in the positions of helix IV with respect to
helix III mostly arise from the large variation in / angle.
For example, some structures (S100A13, S100A1) have /
angles about 608 more positive and others (S100A11,
S100P) have / angles about 608 more negative than
found in human apo-S100B. The VGM plot in Figure 4
fixes helix III along the z-axis and gives the impression
that the N-termini of helix IV, closest to the calcium-
binding loop of EF2 takes on a range of positions with
respect to helix III. Superposition of the apo-S100 pro-
teins, however, shows that it is the C-terminus of helix
III that occupies a variety of locations.
Despite the similarity in y angles observed for the pro-
teins it is tempting to suggest that the range of / angles
observed for the EF2 calcium-binding site might arise
from different structural subfamilies in the apo-S100 pro-
teins. In turn, the different interhelical relationships
between helices III and IV might be important for the
extent that these helices change conformation upon cal-
cium binding. Several factors indicate this is not the case.
For example, statistical analysis of the pairwise superposi-
tions of helices III and IV in EF2 for human apo-S100B
and all other apo-S100 proteins shows a Gaussian distri-
bution [Fig. 5(B)] that is satisfied at the 95% confidence
level and passes several normality tests (with the excep-
tion of S100P). Further, the median of this superposition
(1.85 A˚ ) falls near the midpoint for a Gaussian distribu-
tion of the helices (1.88 Æ 0.73 A˚ ). This result is similar
to that observed for helices I, II and IV (1.72 Æ 0.48 A˚ )
or all four helices (1.84 Æ 0.63 A˚ ) shown in Figure 5.
Using the pairwise rmsd comparisons (see Fig. 5) as a
guide in combination with the VGM analysis (see Fig. 4),
it is apparent the large structural differences arise due to
the displacement of the N-termini of helix IV with
respect to the C-termini of helix III. For example, human
apo-S100B and S100A1 have a relatively small rmsd
Figure 4
Helix orientation for human apo-S100B and other S100 proteins obtained using Vector Geometry Mapping.29,68 The figure shows helices I and II (A, B) from calcium-
binding site EF1 and helices III and IV (C, D) from calcium-binding site EF2. In both cases the incoming helix of each structure is aligned along the z-axis with respect
to the N-terminal helix of EF1 in calmodulin. The exiting helices (II in EF1 and IV in EF2) are shown for the calcium-free proteins calmodulin (grey), human S100B
(red), bovine S100B (brown), rat S100B (maroon), S100A1 (cyan), S100A3 (green), S100A4 (magenta), human S100A6 (blue), rabbit S100A6 (light blue), S100A11
(dark purple), S100A13 (orange), and S100P (yellow). The plots show the helices from calcium-binding site EF1 as viewed down the 1z axis (A, C) and rotated $908
about y and $458 about z (B, D). The directions of the helices are denoted by N?C either along the z-axis or the exiting helix (B, D).
S. Malik et al.
36 PROTEINS
when helices III and IV are superimposed (1.30 A˚ ) yet
the N-termini of helix IV are separated by about 6 A˚ .
This is evident in [Fig. 4(C,D)] where the C-termini of
helix IV from human apo-S100B (red) and apo-S100A1
(cyan) are close together, but their N-termini are more
divergent.
The comparison of the interhelical y and / angles of
the apo-S100B structures shows that human apo-S100B
lies very close to the midpoint of tip and horizontal
angles for both EF1 and EF2 calcium-binding sites. Anal-
ysis of the angular ranges and pairwise backbone rmsd
indicates that all apo-S100 structures to date, with the
exception of S100P (y angle for EF2), fall within a single
Gaussian distribution although the ranges of the horizon-
tal angle (/) for EF2 is nearly sevenfold larger for EF2
than for EF1.
Helix III shows the broadest range
of accessible surface areas in the
S100 proteins
The VGM results indicate there is a significant range
of conformations that helices III and IV adopt with
respect to each other within the S100 family. In order to
pinpoint a rationale for this, the accessible surface area
for human apo-S100B, presented here, and other cal-
cium-free S100 proteins was analyzed. The fraction acces-
sible surface area for human apo-S100B reveals a large
number of residues in helix I (L3, A6, M7, A9, L10, I11,
V13, and F14) and helix IV (F70, F73, M74, F76, V77,
A78, V80, and T81) that have >80% of their side chains
buried (see Fig. 6). Most of these residues form the
hydrophobic dimer interface in apo-S100B derived from
the near perpendicular arrangement of helices I and I0
Figure 5
Comparison of apo-S100 structures showing the distribution of structures as a function of backbone rmsd between helices. The figure shows (A) helices I and II, (B)
helices III and IV, (C) helices I, II, and IV, and (D) sum of all helices for human apo-S100B and the other ten apo-S100 structures described in Table II. The helices used
were L3-Y17 (helix I), K29-E39 (helix II), Q50-L60 (helix III), and F70-V80 (helix IV) in human apo-S100B and the corresponding regions in other S100 proteins based
on alignment using T-Coffee.81 For NMR structures, the most representative structure as listed in the PDB file was used. In each case the rmsd values between all
possible pairs of structures (55 comparisons) were tabulated and binned (0.2 A˚ bins). The number of occurrences was plotted against the bin center. Each graph shows the
best-fit Gaussian curve to the binned rmsd data centered at (A) 1.46 Æ 0.56 A˚, (B) 1.88 Æ 0.73 A˚, (C) 1.72 Æ 0.48 A˚, and (D) 1.84 Æ 0.63 A˚. The structures for apo-
S100A13 and apo-S100P were removed from datasets (A) and (B–D), respectively based on ANOVA statistics.
Conformational Distribution for Apo-S100 Proteins
PROTEINS 37
with IV and IV’. For example, L3 in helix I interacts with
L10 and V13 in helix I0
. In helix IV, residue F73 interacts
with L3 of helix I, and L10 and F14 of helix I0
. Compari-
son with other apo-S100 structures reveals this pattern of
buried hydrophobic residues is nearly perfectly preserved
[Fig. 6(A,D)]. An examination of the S100 sequences
along with this analysis indicates the dimerization motif
for these S100 proteins is LXX[A/C/S][M/L/I/V]XX[M/L/
I/V][I/V]X[V/I/T]F (residues 3-14; helix I) and FXE[F/
Y][V/I/L/M]X[L/F][V/L/I][A/G/S]X[V/L/I][T/A] (residues
70–81; helix IV). The excellent agreement between the
buried surface area and sequence conservation of the apo-
S100 structures is in accord with 15
N relaxation77 and am-
ide proton exchange experiments48 for apo-S100B. In par-
Figure 6
Comparison of side chain fractional accessible surface area (FASA) for residues in human apo-S100B and other apo-S100 structures. Residues in (A) helix I, (B) helix II,
(C) helix III and (IV) helix IV are plotted as a function of fraction accessible surface area by each side chain. The numbering scheme used is that of human apo-S100B
and other proteins were aligned based on sequence using T-Coffee.86 The proteins shown are human S100B (n), bovine S100B (~), rat S100B (!), S100A1 (^),
S100A3 (l), S100A4 (h), human S100A6 (!), rabbit S100A6 (~), S100A11 (^), S100A13 (3) and S100P (*). All fractional accessible surface areas were calculated
from available PDB coordinates using the program VADAR.67 For NMR structures, the most representative structure as listed in the PDB file was used.
S. Malik et al.
38 PROTEINS
ticular, helices I and IV had protection factors that were
similar to free energies for the unfolding of apo-S100B
indicating the dimer interface, maintained by the helix I
and IV motifs is a major contributor towards the stability
of all S100 proteins.
Extending this analysis to helix II shows the hydropho-
bic portions of residues K29, L32, L35, and I36 are all
buried >80% in human apo-S100B. As with helices I
and IV these positions in helix II are highly conserved in
the S100 sequences. The large variation in accessible sur-
face area at the N-terminus of helix II is also accompa-
nied by decreased protection from amide exchange.48 In
helix III, the shortest of the helices in apo-S100B, a large
variation in the accessible surface area exists between
structures. Only two residues (V56, L60) show a consist-
ent pattern of burial between human apo-S100B and the
other S100 proteins and these are both accompanied by
high conservation at these positions. Other residues are
less well conserved and show poorer patterns of accessi-
ble surface area. For example, V52 in the apo-S100B
structures is nearly completely buried while residues in
S100A6 (E52) and S100A3 (D54) are more exposed con-
sistent with their significant differences in side chain po-
larity. Notably, position 57 in helix III, where methionine
is highly conserved, shows an inconsistent pattern. In
human apo-S100B it shows about 30% accessible surface
area, midway between the observed range for this residue
(2%–62%), despite the conserved nature of the methio-
nine residue in all sequences but S100A13 and S100P.
Further differences are seen at the second position in the
calcium-binding loop occupied by N62 in human apo-
S100B, where it is exposed but nearly completely buried
in S100A11 where a leucine residue occupies this posi-
tion. The large variations in buried surface area through-
out helix III and especially for conserved residues (V53,
M57) are most consistent with this helix being loosely
packed. This conclusion is borne out by amide hydrogen
exchange rates for residues in helix III that are 2–3 orders
of magnitude faster than helices I, II, and IV.48 In addi-
tion, although amide exchange rates have not been meas-
ured for rat or bovine apo-S100B, 15
N relaxation experi-
ments indicate helix III in rat apo-S100B has the lowest
order parameters of the four helices.77 A similar obser-
vation has been made in the N-terminal domain of tro-
ponin-C78,79 where helix C, one of the helices that
undergoes a conformational change upon calcium bind-
ing, has the lowest order parameters of the helices in that
domain.
Function relevance for the S100 proteins
An analysis of the tip and horizontal angles for the
EF1 and EF2 calcium-binding sites in human apo-S100B
and several other apo-S100 proteins reveals a range of
orientations are adopted by EF2 with a mean near that
determined for human apo-S100B. It was also shown
that helix III is clearly the most loosely packed of the
four helices in all structures of apo-S100 proteins. This
data indicates that helix III is likely a flexible helix able
to sample a range of orientations with respect to helix
IV. Is this important for the function of the S100 pro-
teins? As shown in Figure 7, the broad range of tip and
horizontal angles displayed in the apo-S100 structures is
not exhibited in the calcium-bound forms. As with the
apo-proteins the structure of the calcium-bound form of
human S100B has tip and horizontal angles for EF2 (818,
100.18) that are representative of the average angles of
67.5 Æ 9.5 and 100.3 Æ 8.88 for the calcium-bound
structures. For human S100B, this indicates that the pro-
tein undergoes changes of about 558 and 1148 in the tip
and horizontal angles respectively in response to calcium
binding. This opening of the horizontal angle is 2.5–3.0
times larger than observed for calmodulin82 or
troponin-C.83,84 This structural change is responsible
for the shallow binding surface noted in the calcium-
bound forms of the S100 proteins compared to a nar-
rower cleft found in calmodulin and troponin-C.85,86 It
is interesting that the calcium-bound S100 structures
compared (4 NMR, 6 X-ray structures) display a much
narrower range of horizontal angles (100.3 Æ 8.88) clearly
indicating a more homogeneous group of structures
compared to their apo forms (Table II). This would indi-
cate that calcium binding might act to restrict the freedom
of helix III, presumably through an ordering of the EF2 cal-
cium-binding site. For S100B, amide exchange experiments
Figure 7
Helix reorientation for the EF2 calcium-binding site in S100 proteins. The
relative distribution for each S100 protein is plotted versus the horizontal angle
(/) calculated using Vector Geometry Mapping.29,68 NMR structures are
represented as a Gaussian curve centered at the mean and having a width at
half-height corresponding to Æ the standard deviation. X-ray structures are
shown as vertical sticks. The structures shown are; human S100B (red), bovine
S100B (orange), rat S100B (maroon), S100A1 (green), S100A3 (ochre), S100A4
(blue), human S100A6 (sky blue), rabbit S100A6 (dark green), S100A11
(magenta), S100A12 (pink) S100A13 (rose), and S100P (black).
Conformational Distribution for Apo-S100 Proteins
PROTEINS 39
have shown that the amide rates of exchange in this cal-
cium-binding site are slowed by 1–2 orders of magni-
tude. However, amide exchange rates for helix III remain
the fastest of the helices in S100B indicating this helix is
still exposed.48 Still, Figure 7 provides evidence that cal-
cium binding to the S100 proteins enables EF2 to move
from a broad range of conformations to a tighter more
compressed distribution.
CONCLUSIONS
Based on amide chemical shift changes and residual
dipolar changes the three-dimensional structure of
human apo-S100B does not appear to undergo signifi-
cant structural changes between 35 and 108C. A compari-
son of the three dimensional structure of human apo-
S100B with those of rat and bovine apo-S100B indicates
the human form has similar helix orientations and preci-
sion as the rat form of the protein but has calcium-bind-
ing loops more similar to those obtained for the bovine
structure. Using human apo-S100B as a template the
interhelical angles for EF-hands EF1 and EF2 were com-
pared with other apo-S100 protein structures. The
arrangements of EF1 (y 5 72 Æ 4, / 5 98 Æ 8) and
EF2 (y 5 25 Æ 2, / 5 2146 Æ 8) were found to be
near the mean for all apo-S100 protein structures deter-
mined to date. Although a broad distribution of the EF2
helical angles exist for the apo-S100 proteins, we have
shown that these structures assume a normal Gaussian
distribution about this conformation indicative of a sin-
gle structural population. Calcium binding alters the con-
formation of the EF2 helices (III, IV) but more signifi-
cantly leads to a tighter less variable arrangement of the
helices.
ACKNOWLEDGMENTS
The authors would like to thanks Kathryn Barber
(UWO) for her technical support. We are grateful to
Lewis Kay (University of Toronto) for providing pulse
sequences, Frank Delaglio for NMRPipe and DYNAMO,
and Bruce Johnson for NMRView. We would like to
thank the Canadian National High Field NMR Centre
(NANUC) for their assistance and use of the facilities.
Operation of NANUC is funded by the Canadian Insti-
tutes of Health Research, the Natural Science and Engi-
neering Research Council of Canada and the University
of Alberta.
REFERENCES
1. Santamaria-Kisiel L, Rintala-Dempsey AC, Shaw GS. Calcium-de-
pendent and -independent interactions of the S100 protein family.
Biochem J 2006;396:201–214.
2. Schafer BW, Wicki R, Englekamp D, Mattei MG, Heizmann CW.
Isolation of a YAC clone covering a cluster of nine S100 genes
on human chromosome 1q21: rationale for a new nomenclature of
the S100 calcium-binding protein family. Genomics 1995;25:638–
643.
3. Vogl T, Ludwig S, Goebeler M, Strey A, Thorey IS, Reichelt R, Foell
D, Gerke V, Manitz MP, Nacken W, Werner S, Sorg C, Roth J.
MRP8 and MRP14 control microtubule reorganization during
transendothelial migration of phagocytes. Blood 2004;104:4260–
4268.
4. Frizzo JK, Tramontina AC, Tramontina F, Gottfried C, Leal RB,
Donato R, Goncalves CA. Involvement of the S100B in cAMP-
induced cytoskeleton remodeling in astrocytes: a study using
TRTK-12 in digitonin-permeabilized cells. Cell Mol Neurobiol
2004;24:833–840.
5. Bianchi R, Garbuglia M, Verzini M, Giambanco I, Ivanenkov VV,
Dimlich RV, Jamieson GA, Jr, Donato R. S-100 (alpha and beta)
binding peptide (TRTK-12) blocks S-100/GFAP interaction: identi-
fication of a putative S-100 target epitope within the head domain
of GFAP. Biochim Biophys Acta 1996;1313:258–267.
6. Davey GE, Murmann P, Hoechli M, Tanaka T, Heizmann CW. Cal-
cium-dependent translocation of S100A11 requires tubulin fila-
ments. Biochim Biophys Acta 2000;1498:220–232.
7. Garbuglia M, Verzini M, Rustandi RR, Osterloh D, Weber DJ,
Gerke V, Donato R. Role of the C-terminal extension in the interac-
tion of S100A1 with GFAP, tubulin, the S100A1- and S100B-inhibi-
tory peptide. TRTK-12, and a peptide derived from p53, and the
S100A1 inhibitory effect on GFAP polymerization. Biochem Bio-
phys Res Comm 1999;254:36–41.
8. Deora AB, Kreitzer G, Jacovina AT, Hajjar KA. An annexin 2 phos-
phorylation switch mediates p11-dependent translocation of
annexin 2 to the cell surface. J Biol Chem 2004;279:43411–43418.
9. Bichsel SJ, Tamaskovic R, Stegert MR, Hemmings BA. Mechanism
of activation of NDR (nuclear Dbf2-related) protein kinase by the
hMOB1 protein. J Biol Chem 2004;279:35228–35235.
10. Wilder PT, Lin J, Bair CL, Charpentier TH, Yang D, Liriano M, Var-
ney KM, Lee A, Oppenheim AB, Adhya S, Carrier F, Weber DJ. Rec-
ognition of the tumor suppressor protein p53 and other protein
targets by the calcium-binding protein S100B. Biochim Biophys
Acta 2006;1763:1284–1297.
11. Sheu F-S, Huang FL, Huang K-P. Differential responses of protein
kinase C substrates (MARCKS, neuromodulin and neurogranin)
phosphorylation to calmodulin and S100. Arch Biochem Biophys
1995;316:335–342.
12. Baudier J, Bergeret E, Bertacchi N, Weintraub H, Gagnon J, Garin
J. Interactions of myogenic bHLH transcription factors with cal-
cium-binding calmodulin and S100a (alpha alpha) proteins. Bio-
chemistry 1995;34:7834–7846.
13. Baudier J, Bronner C, Kligman D, Cole RD. Protein kinase C sub-
strates from bovine brain. J Biol Chem 1989;264:1824–1828.
14. Hatakeyama T, Okada M, Shimamoto S, Kubota Y, Kobayashi R.
Identification of intracellular target proteins of the calcium-signal-
ing protein S100A12. Eur J Biochem 2004;271:3765–3775.
15. Zimmer DB, Van Eldik LJ. Identification of a molecular target for
the calcium-modulated protein S100. Fructose-1,6-bisphosphate al-
dolase. J Biol Chem 1986;261:11424–11428.
16. Landar A, Caddell G, Chessher J, Zimmer DB. Identification of an
S100A1/S100B target protein: phosphoglucomutase. Cell Calcium
1996;20:279–285.
17. Lewit-Bentley A, Rety S, Sopkova-de OliveiraSantos J, Gerke V.
S100-annexin complexes: some insights from structural studies. Cell
Biol Int 2000;24:799–802.
18. Seemann J, Weber K, Gerke V. Annexin I targets S100C to early
endosomes. FEBS Lett 1997;413:185–190.
19. Gerke V. Tyrosine protein kinase substrate p36: a member of the
annexin family of Ca21
/phospholipid-binding proteins. Cell Motil
Cytoskeleton 1989;14:449–454.
20. Arcuri C, Giambanco I, Bianchi R, Donato R. Annexin V, annexin
VI, S100A1 and S100B in developing and adult avian skeletal
muscles. Neuroscience 2002;109:371–388.
S. Malik et al.
40 PROTEINS
21. Chang N, Sutherland C, Hesse E, Winkfein R, Wiehler WB, Pho M,
Veillette C, Li S, Wilson DP, Kiss E, Walsh MP. Identification of a
novel interaction between the Ca(21)-binding protein S100A11
and the Ca(21)- and phospholipid-binding protein annexin A6.
Am J Physiol Cell Physiol 2007;292:C1417–C1430.
22. Nowotny M, Spiechowicz M, Jastrzebska B, Filipek A, Kitagawa K,
Kuznicki J. Calcium-regulated interaction of Sgt1 with S100A6 (cal-
cyclin) and other S100 proteins. J Biol Chem 2003;278:26923–
26928.
23. Filipek A, Jastrzebska B, Nowotny M, Kuznicki J. CacyBP/SIP, a cal-
cyclin and Siah-1-interacting protein, binds EF-hand proteins of the
S100 family. J Biol Chem 2002;277:28848–28852.
24. Fujii T, Hiromori T, Hamamoto M, Suzuki T. Interaction of
chicken gizzard smooth muscle calponin with brain microtubules.
J Biochem (Tokyo) 1997;122:344–351.
25. Baudier J, Delphin C, Grunwald D, Khochbin S, Lawrence JJ. Char-
acterization of the tumor suppressor protein p53 as a protein kinase
C substrate and a S100b-binding protein. Proc Natl Acad Sci U S A
1992;89:11627–11631.
26. Lin J, Blake M, Tang C, Zimmer D, Rustandi RR, Weber DJ, Carrier
F. Inhibition of p53 transcriptional activity by the S100B calcium-
binding protein. J Biol Chem 2001;276:35037–35041.
27. Scotto C, Deloulme JC, Rousseau D, Chambaz E, Baudier J. Cal-
cium and S100B regulation of p53-dependent cell growth arrest and
apoptosis. Mol Cell Biol 1998;18:4272–4281.
28. Yu WH, Fraser PE. S100B interaction with Tau is promoted by zinc
and inhibited by hyperphosphorylation in Alzheimer’s disease.
J Neurosci 2001;21:2240–2246.
29. Yap KL, Ames JB, Swindells MB, Ikura M. Diversity of conforma-
tional states and changes within the EF-hand protein superfamily.
Proteins 1999;37:499–507.
30. Ikura M. Calcium binding and conformational response in EF-hand
proteins. Trends Biochem 1996;21:14–17.
31. Marshak DR, Pesce SA, Stanley LC, Griffin WST. Increased S100b
neurotrophic activity in Alzheimer’s disease temporal lobe. Neuro-
biol Aging 1991;13:1–7.
32. Boom A, Pochet R, Authelet M, Pradier L, Borghgraef P, Van
Leuven F, Heizmann CW, Brion JP. Astrocytic calcium/zinc binding
protein S100A6 over expression in Alzheimer’s disease and in PS1/
APP transgenic mice models. Biochim Biophys Acta 2004;1742:161–
168.
33. Yui S, Nakatani Y, Mikami M. Calprotectin (S100A8/S100A9), an
inflammatory protein complex from neutrophils with a broad apo-
ptosis-inducing activity. Biol Pharm Bull 2003;26:753–760.
34. Odink K, Cerletti N, Bruggen J, Clerc RG, Tarcsay L, Zwadlo G,
Gerhards G, Schlegel R, Sorg C. Two calcium-binding proteins in
infiltrate macrophages of rheumatoid arthritis. Nature 1987;330:80–
82.
35. Deloulme JC, Gentil BJ, Baudier J. Monitoring of S100 homodime-
rization and heterodimeric interactions by the yeast two-hybrid sys-
tem. Microsc Res Tech 2003;60:560–568.
36. Deloulme JC, Assard N, Mbele GO, Mangin C, Kuwano R, Baudier
J. S100A6 and S100A11 are specific targets of the calcium- and
zinc-binding S100B protein in vivo. J Biol Chem 2000;275:35302–
35310.
37. Potts BCM, Smith J, Akke M, Macke TJ, Okazaki K, Hidaka H,
Case DA, Chazin WJ. The structure of calcyclin reveals a novel
homodimeric fold for S100 Ca21
-binding proteins. Nature Struct
Biol 1995;2:790–796.
38. Rety S, Sopkova J, Renouard M, Osterloh D, Gerke V, Tabaries S,
Russo-Marie R, Lewit-Bentley A. The crystal structure of a complex
of p11 with the annexin II N-terminal peptide. Nature Struct Biol
1999;6:89–95.
39. Rety S, Osterloh D, Arie JP, Tabaries S, Seeman J, Russo-Marie F,
Gerke V, Lewit-Bentley A. Structural basis of the Ca21
-dependent
association between S100C (S100A11) and its target, the N-terminal
part of annexin I. Structure 2000;8:175–184.
40. Drohat AC, Tjandra N, Baldisseri DM, Weber DJ. The use of dipo-
lar couplings for determining the solution structure of rat apo-
S100B(bb). Protein Sci 1999;8:800–809.
41. Drohat AC, Baldisseri DM, Rustandi RR, Weber DJ. Solution structure
of calcium-bound rat S100B(betabeta) as determined by nuclear mag-
netic resonance spectroscopy. Biochemistry 1998;37:2729–2740.
42. Smith SP, Shaw GS. A novel calcium-sensitive switch revealed by
the structure of human S100B in the calcium-bound form. Struc-
ture 1998;6:211–222.
43. Kilby PM, Van Eldik LJ, Roberts GCK. The solution structure of
the bovine S100b protein dimer in the calcium-free state. Structure
1996;4:1041–1052.
44. Maler L, Potts BC, Chazin WJ. High resolution solution structure
of apo calcyclin and structural variations in the S100 family of cal-
cium-binding proteins. J Biomol NMR 1999;13:233–247.
45. Sastry M, Ketchem RR, Crescenzi O, Weber C, Lubienski MJ,
Hidaka H, Chazin WJ. The three-dimensional structure of Ca21
-
bound calcyclin: implications for Ca21
-signal transduction by S100
proteins. Structure 1998;6:223–231.
46. Otterbein LR, Kordowska J, Witte-Hoffmann C, Wang C-L, Domi-
nguez R. Crystal structures of S100A6 in the Ca21
-free and Ca21
-
bound states: the calcium sensor mechanism of S100 proteins
revealed at atomic resolution. Structure 2002;10:557–567.
47. Dempsey AC, Walsh MP, Shaw GS. Unmasking the annexin I inter-
action from the structure of Apo-S100A11. Structure (Camb)
2003;11:887–897.
48. Marlatt NM, Shaw GS. Amide exchange shows calcium-induced
conformational changes are transmitted to the dimer interface of
S100B. Biochemistry 2007;46:7478–7487.
49. Smith SP, Barber KR, Dunn SD, Shaw GS. Structural influence of
cation binding to recombinant human brain S100b: evidence for
calcium-induced exposure of a hydrophobic surface. Biochemistry
1996;35:8805–8814.
50. Bax A, Ikura M. An efficient 3D NMR technique for correlating the
proton and 15N backbone amide resonances with the a-carbon of
the preceding residue in uniformly 15N/13C enriched proteins.
J Biomol NMR 1991;1:99–104.
51. Wittekind M, Mueller L. HNCACB, A high sensitivity 3D NMR
experiment to correlate amide proton and nitrogen resonances with
the a-carbon and b-carbon resonances in proteins. J Magn Reson
Series B 1993;101:171–180.
52. Grzesiek S, Bax A. An efficient experiment for sequential backbone
assignment of medium-sized isotopically enriched proteins. J Magn
Reson 1992;99:201–207.
53. Grzesiek S, Anglister J, Bax A. Correlation of backbone amide and
aliphatic side-chain resonances in 13
C/15
N-enriched proteins by iso-
tropic mixing of 13
C magnetization. J Magn Reson Series B
1993;101:114–119.
54. Kay LE, Xu G, Singer AU, Muhandiram DR, Forman-Kay JD. A
gradient-enhanced HCCH-TOCSY experiment for recording side-
chain 1
H and 13
C correlations in H2O samples of proteins. J Magn
Reson 1993;101:333–337.
55. Yamazaki T, Forman-Kay JD, Kay LE. Two-dimensional NMR
experiments for correlating 13
Cb and 1
Hd/e chemical shifts of aro-
matic residues in 13
C-labeled proteins via scalar couplings. J Am
Chem Soc 1993;115:11054–11055.
56. Zwahlen C, Legault P, Vincent SJF, Greenblatt J, Konrat R, Kay
LE. Methods for measurement of intermolecular NOEs by multi-
nuclear NMR spectroscopy: application to a bateriophage k N-
Peptide/boxB RNA complex. J Am Chem Soc 1997;119:6711–
6721.
57. Ottiger M, Bax A. Bicelle-based liquid crystals for NMR-measure-
ment of dipolar couplings at acidic and basic pH values. J Biomol
NMR 1999;13:187–191.
58. Tjandra N, Bax A. Direct measurement of distances and angles in
biomolecules by NMR in a dilute liquid crystalline medium. Science
1997;278:1111–1114.
Conformational Distribution for Apo-S100 Proteins
PROTEINS 41
59. Hansen MR, Mueller L, Pardi A. Tunable alignment of macromole-
cules by filamentous phage yields dipolar coupling interactions. Nat
Struct Biol 1998;5:1065–1074.
60. Delaglio F, Grzesiek S, Vuister GW, Zhu G, Pfeifer J, Bax A.
NMRPipe: a multidimensional spectral processing system based on
UNIX pipes. J Biomol NMR 1995;6:277–293.
61. Johnson BA, Belvins RA. NMRView: a computer program for the
visualization and analysis of NMR data. J Biomol NMR 1994;4:
603–614.
62. Brunger AT, Adams PD, Clore GM, DeLano WL, Gros P, Grosse-
Kunstleve RW, Jiang JS, Kuszewski J, Nilges M, Pannu NS, Read RJ,
Rice LM, Simonson T, Warren GL. Crystallography & NMR system:
a new software suite for macromolecular structure determination.
Acta Crystallogr D Biol Crystallogr 1998;54(Part 5):905–921.
63. Gagne SM, Tsuda S, Li MX, Chandra M, Smillie LB, Sykes BD.
Quantification of the calcium-induced secondary structural changes
in the regulatory domain of troponin-C. Protein Sci 1994;3:1961–
1974.
64. Cornilescu G, Delagio F, Bax A. Protein backbone angle restraints
from searching a database for chemical shift and sequence homol-
ogy. J Biomol NMR 1999;13:289–302.
65. Nederveen AJ, Doreleijers JF, Vranken W, Miller Z, Spronk CA,
Nabuurs SB, Guentert P, Livny M, Markley JL, Nilges M, Ulrich EL,
Kaptein R, Bonvin JJ. RECOORD: a recalculated coordinate data-
base of 5001 proteins from that PDB using restraints from that
BioMagResBank. Proteins 2005;59:662–672.
66. Koradi R, Billeter M, Wuthrich K. MOLMOL: a program for dis-
play and analysis of macromolecular structures. J Mol Graphics
1996;14:51–55.
67. Willard L, Ranjan A, Zhang H, Monzavi H, Boyko RF, Sykes BD,
Wishart DS. VADAR: a web server for quantitative evaluation of
protein structure quality. Nucleic Acids Res 2003;31:3316–3319.
68. Yap KL, Ames JB, Swindells MB, Ikura M. Vector geometry mapping.
A method to characterize the conformation of helix-loop-helix cal-
cium-binding proteins. Methods Mol Biol 2002;173:317–324.
69. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM,
Meng EC, Ferrin TE. UCSF Chimera—a visualization system for ex-
ploratory research and analysis. J Comput Chem 2004;25:1605–1612.
70. Lee YC, Volk DE, Thiviyanathan V, Kleerekoper Q, Gribenko AV,
Zhang S, Gorenstein DG, Makhatadze GI, Luxon BA. NMR
structure of the Apo-S100P protein. J Biomol NMR 2004;29:399–
402.
71. Fritz G, Mittl PRE, Vasak M, Grutters MG, Heizmann CW. The
crystal structure of metal-free human EF-hand protein S100A3 at
1.7-A resolution. J Biol Chem 2002;277:33092–33098.
72. Tsuda S, Miura A, Gagne SM, Spyracopoulos L, Sykes BD. Low-
temperature-induced structural changes in the Apo regulatory do-
main of skeletal muscle troponin C. Biochemistry 1999;38:5693–
5700.
73. Barber KR, McClintock KA, Jamieson GA, Jr, Dimlich RV, Shaw
GS. Specificity and Zn21
enhancement of the S100B binding epi-
tope TRTK-12. J Biol Chem 1999;274:1502–1508.
74. Smith SP, Shaw GS. Assignment and secondary structure of cal-
cium-bound human S100B. J Biomol NMR 1997;10:77–88.
75. Baxter NJ, Williamson MP. Temperature dependence of 1H chemi-
cal shifts in proteins. J Biomol NMR 1997;9:359–369.
76. Smith SP, Barber KR, Shaw GS. Identification and structural influ-
ence of a differentially modified N-terminal methionine in human
S100b. Protein Sci 1997;6:1110–1113.
77. Inman KG, Baldisseri DM, Miller KE, Weber DJ. Backbone dynam-
ics of the calcium-signaling protein apo-S100B as determined by
15N NMR relaxation. Biochemistry 2001;40:3439–3448.
78. Gagne SM, Tsuda S, Spyrocopoulos L, Kay LE, Sykes BD. Backbone
and methyl dynamics of the regulatory domain of troponin C: ani-
sotropic rotational diffusion and contribution of conformational
entropy to calcium affinity. J Mol Biol 1998;278:667–686.
79. Spyracopoulos L, Lavigne P, Crump MP, Gagne SM, Kay CM, Sykes
BD. Temperature dependence of dynamics and thermodynamics of
the regulatory domain of human cardiac troponin C. Biochemistry
2001;40:12541–12551.
80. Babini E, Bertini I, Capozzi F, Luchinat C, Quattrone A, Turano M.
Principal component analysis of the conformational freedom within
the EF-hand superfamily. J Proteome Res 2005;4:1961–1971.
81. Notredame C, Higgins DG, Heringa J. T-Coffee: a novel method for
fast and accurate multiple sequence alignment. J Mol Biol 2000;302:
205–217.
82. Zhang M, Tanaka T, Ikura M. Calcium-induced conformational
transition revealed by the solution structure of apo calmodulin.
Nature Struct Biol 1995;2:758–767.
83. Gagne SM, Li MX, McKay RT, Sykes BD. The NMR angle on tro-
ponin C. Biochem Cell Biol 1998;76:302–312.
84. Gagne SM, Tsuda S, Li MX, Smillie LB, Sykes BD. Structures of the
troponin C regulatory domains in the apo and calcium-saturated
states. Nat Struct Biol 1995;2:784–789.
85. Bhattacharya S, Bunick CG, Chazin WJ. Target selectivity in EF-
hand calcium binding proteins. Biochim Biophys Acta 2004;1742:
69–79.
86. Maler L, Sastry M, Chazin WJ. A structural basis for S100 protein
specificity derived from comparative analysis of apo and Ca21
-cal-
cyclin. J Mol Biol 2002;317:279–290.
S. Malik et al.
42 PROTEINS

More Related Content

What's hot

2012_SFRRi_Integrated Omics
2012_SFRRi_Integrated Omics2012_SFRRi_Integrated Omics
2012_SFRRi_Integrated Omics
Ana Reis
 
Clocl_2077 crystallization- FINAL
Clocl_2077 crystallization- FINALClocl_2077 crystallization- FINAL
Clocl_2077 crystallization- FINAL
ODED MIZRACHI
 
Open stax biology (nonmajors) ch06
Open stax biology (nonmajors) ch06Open stax biology (nonmajors) ch06
Open stax biology (nonmajors) ch06
Lumen Learning
 
Gutell 092.jmb.2004.344.1225
Gutell 092.jmb.2004.344.1225Gutell 092.jmb.2004.344.1225
Gutell 092.jmb.2004.344.1225
Robin Gutell
 

What's hot (13)

2012_SFRRi_Integrated Omics
2012_SFRRi_Integrated Omics2012_SFRRi_Integrated Omics
2012_SFRRi_Integrated Omics
 
Cell Structure and Function
Cell Structure and FunctionCell Structure and Function
Cell Structure and Function
 
2006 O'Leary et al MBC
2006 O'Leary et al  MBC2006 O'Leary et al  MBC
2006 O'Leary et al MBC
 
1
11
1
 
The great basic question of science: Membrane compartment or non-membrane pha...
The great basic question of science: Membrane compartment or non-membrane pha...The great basic question of science: Membrane compartment or non-membrane pha...
The great basic question of science: Membrane compartment or non-membrane pha...
 
Genes Dev
Genes DevGenes Dev
Genes Dev
 
Hydrogenosome in fungi
Hydrogenosome in fungiHydrogenosome in fungi
Hydrogenosome in fungi
 
Origin and function of a metal resistance gene isolated from eukaryotic soil ...
Origin and function of a metal resistance gene isolated from eukaryotic soil ...Origin and function of a metal resistance gene isolated from eukaryotic soil ...
Origin and function of a metal resistance gene isolated from eukaryotic soil ...
 
Clocl_2077 crystallization- FINAL
Clocl_2077 crystallization- FINALClocl_2077 crystallization- FINAL
Clocl_2077 crystallization- FINAL
 
Open stax biology (nonmajors) ch06
Open stax biology (nonmajors) ch06Open stax biology (nonmajors) ch06
Open stax biology (nonmajors) ch06
 
Gutell 092.jmb.2004.344.1225
Gutell 092.jmb.2004.344.1225Gutell 092.jmb.2004.344.1225
Gutell 092.jmb.2004.344.1225
 
853
853853
853
 
Asea component verification_document
Asea component verification_documentAsea component verification_document
Asea component verification_document
 

Similar to APO-S100B

tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...
tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...
tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...
saheli halder
 
5. all published paper
5. all published paper5. all published paper
5. all published paper
Halei Zhai
 
ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016
Alex Kiselyov
 
Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...
Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...
Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...
Seamus Quaid
 
boe-6-10-4098
boe-6-10-4098boe-6-10-4098
boe-6-10-4098
Gen Vigil
 
Calcium signaling in_plant_cell_organelles_delimited_by_a_double_membrane
Calcium signaling in_plant_cell_organelles_delimited_by_a_double_membraneCalcium signaling in_plant_cell_organelles_delimited_by_a_double_membrane
Calcium signaling in_plant_cell_organelles_delimited_by_a_double_membrane
Diego Urrutia
 
Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...
Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...
Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...
Igor Putrenko
 

Similar to APO-S100B (20)

tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...
tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...
tuning the pH Response of i-Motif DNA Oligonucleotides_Lannes_et_al-2015-Chem...
 
5. all published paper
5. all published paper5. all published paper
5. all published paper
 
Potassium (K+) channels
Potassium (K+) channelsPotassium (K+) channels
Potassium (K+) channels
 
Cyclic conformation and nucleic acid sugar puckering
Cyclic conformation and nucleic acid sugar puckeringCyclic conformation and nucleic acid sugar puckering
Cyclic conformation and nucleic acid sugar puckering
 
journal.pone.0128622
journal.pone.0128622journal.pone.0128622
journal.pone.0128622
 
ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016ASK_Modeling_PaperBMC2016
ASK_Modeling_PaperBMC2016
 
Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...
Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...
Investigating the effects of oxysterols 4Beta-OH and 7Beta-OH on calcium sign...
 
cmb(1) calcium.pptx
cmb(1) calcium.pptxcmb(1) calcium.pptx
cmb(1) calcium.pptx
 
C4DT02257B
C4DT02257BC4DT02257B
C4DT02257B
 
S100 protein you need to know from Feiyuebio.pdf
S100 protein you need to know from Feiyuebio.pdfS100 protein you need to know from Feiyuebio.pdf
S100 protein you need to know from Feiyuebio.pdf
 
Feiyuebio S100 protein.pdf
Feiyuebio S100 protein.pdfFeiyuebio S100 protein.pdf
Feiyuebio S100 protein.pdf
 
4
44
4
 
boe-6-10-4098
boe-6-10-4098boe-6-10-4098
boe-6-10-4098
 
Calcium signaling in_plant_cell_organelles_delimited_by_a_double_membrane
Calcium signaling in_plant_cell_organelles_delimited_by_a_double_membraneCalcium signaling in_plant_cell_organelles_delimited_by_a_double_membrane
Calcium signaling in_plant_cell_organelles_delimited_by_a_double_membrane
 
Characterization of the lipid binding properties of otoferlin reveals specifi...
Characterization of the lipid binding properties of otoferlin reveals specifi...Characterization of the lipid binding properties of otoferlin reveals specifi...
Characterization of the lipid binding properties of otoferlin reveals specifi...
 
M1-Biophysics
M1-BiophysicsM1-Biophysics
M1-Biophysics
 
Renal physiology
Renal physiologyRenal physiology
Renal physiology
 
s.samanta
s.samantas.samanta
s.samanta
 
Bis-pyrene probes of foldamer conformation in solution and in phospholipid bi...
Bis-pyrene probes of foldamer conformation in solution and in phospholipid bi...Bis-pyrene probes of foldamer conformation in solution and in phospholipid bi...
Bis-pyrene probes of foldamer conformation in solution and in phospholipid bi...
 
Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...
Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...
Modulation of Flash-Induced Photosystem II Fluorescence by Events Occurring a...
 

APO-S100B

  • 1. proteinsSTRUCTURE O FUNCTION O BIOINFORMATICS Analysis of the structure of human apo-S100B at low temperature indicates a unimodal conformational distribution is adopted by calcium-free S100 proteins Shahid Malik,1 Matthew Revington,1 Steven P. Smith,2 and Gary S. Shaw1 * 1 Department of Biochemistry, The University of Western Ontario, London, Ontario, N6A 5C1, Canada 2 Department of Biochemistry, Queen’s University, Kingston, Ontario, K7L 3N6, Canada INTRODUCTION The S100 protein family makes up one of the largest EF-hand groups of proteins, comprising at least 25 members in humans.1,2 These low molecular weight (21–24 kDa) proteins are generally expressed in a cell-specific manner having a variety of biological roles including the regulation of cytoskeletal protein assembly,3–7 control of protein phosphorylation,8–13 and modu- lation of the activities of enzymes such as aldolase14,15 and phosphoglucomutase.16 To date over 90 potential protein targets have been identified for the S100 proteins,1 which include the annexins (A1, A2, A5, A6, A11),17–21 the ubiquitination protein CacyBP,22,23 F-actin,24 p53,25–27 and tau.28 Most of these interactions are controlled through calcium binding and a subse- quent conformational change of the S100 protein, revealing a hydrophobic surface that recognizes the target protein sequence. In general, this hydrophobic binding surface can also be probed through binding of the calcium-bound S100 protein to phenyl sepharose or anilinonaphthalene-8-sulfonic acid (ANS). These probes have proven to be a connecting trademark for the S100 proteins, placing them within the calcium ‘‘sensor’’ group of EF- hand proteins that also includes the muscle protein troponin-C and the multifunctional protein calmodulin.29,30 Further interest in the S100 members has evolved because some of these proteins, such as S100B and S100A8/S100A9 show elevated expression lev- els in Alzheimer’s disease31,32 and rheumatoid arthritis,33,34 respectively, suggesting they may have roles in these diseases. S100B is one of the best-characterized members of the S100 protein family. As with all other S100 proteins, except S100G (calbindin D9k), S100B can form homodimers, comprising two noncovalently associated 91-residue subunits. Further, yeast two- hybrid and optical biosensor experiments have indicated that Abbreviations: nOe, nuclear overhauser effect; RDC, residual dipolar couplings; rmsd, root mean square deviation. Grant sponsor: Canadian Institutes of Health Research (GSS). *Correspondence to: Gary S. Shaw, Department of Biochemistry, The University of Western Ontario, London, Ontario, N6A 5C1, Canada. E-mail: gshaw1@uwo.ca. Received 30 November 2007; Revised 8 February 2008; Accepted 19 February 2008 Published online 2 April 2008 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/prot.22037 ABSTRACT S100B is one of the best-characterized members of the calcium-signaling S100 protein family. Most S100 proteins are dimeric, with each monomer containing two EF-hand calcium-binding sites (EF1, EF2). S100B and other S100 proteins respond to calcium increases in the cell by coordinating calcium and undergoing a conformational change that allows them to interact with a variety of cellular targets. Although several three dimensional structures of S100 proteins are available in the calcium-free (apo-) state it has been observed that these structures appear to adopt a wide range of conformations in the EF2 site with respect to the positioning of helix III, the helix that undergoes the most dramatic calcium-induced conformational change. In this work, we have determined the struc- ture of human apo-S100B at 108C to examine whether temperature might be responsible for these structural differences. Further, we have used this data, and other available apo-S100 structures, to show that despite the range of interhelical angles adopted in the apo-S100 structures, normal Gaussian distributions about the mean angles found in the structure of human apo-S100B are observed. This finding, only obvious from the analysis of all avail- able apo-S100 proteins, provides direct structural evi- dence that helix III is a loosely packed helix. This is likely a necessary functional property of the S100 pro- teins that facilitates the calcium-induced conforma- tional change of helix III. In contrast, the calcium- bound structures of the S100 proteins show signifi- cantly smaller variability in the interhelical angles. This shows that calcium binding to the S100 proteins causes not only a conformational change but results in a tighter distribution of helices within the EF2 cal- cium binding site required for target protein interac- tions. Proteins 2008; 73:28–42. VVC 2008 Wiley-Liss, Inc. Key words: calcium-binding proteins; structure; nmr. 28 PROTEINS VVC 2008 WILEY-LISS, INC.
  • 2. S100B can form heterodimers in the cell with S100A1, S100A6, and S100A11.35,36 Each S100B monomer com- prises a globular domain containing two EF-hand cal- cium-binding motifs. Sequence comparison of the S100 proteins reveals the presence of a basic N-terminal pseudo (C) calcium-binding site, comprising 14 residues (site I) between helices I and II (EF1), and an acidic 12- residue canonical calcium-binding site in the C-terminus (site II) flanked by helices III and IV (EF2). The region joining the two calcium-binding sites (linker), and the extreme N- and C-terminal regions of the S100 family members are more divergent in their sequences.37 Three- dimensional structures of S100A1038 and S100A1139 show that residues within these latter three regions are important for the interactions with the annexin proteins that are thought to be important in membrane vesicula- tion processes. Three-dimensional structures of several S100 proteins in the calcium-free (apo) and calcium-bound states, including S100B,40–43 S100A644–46 and S100A11,39,47 shows that they undergo a significant conformational change on binding calcium, exposing several hydrophobic residues that are required for interacting with specific bi- ological target molecules. For example, the conforma- tional change in S100A11 indicates a 408 reorientation of helix III occurs on calcium binding leading to a more open or exposed conformation. In other S100 proteins, such as S100B and S100A6, the calcium-induced confor- mational change ranges between $608 and $908. These observations might indicate that the calcium-induced structural change is different for each S100 protein. Alter- natively, the structures could reflect a wide range of con- formational space that can be sampled by helix III in ei- ther the apo- or calcium-bound structures. For example, the unstable nature of helix III in apo-S100B is consistent with amide protons in this helix that exchange up to three orders of magnitude faster than those in helices I, II, or IV.48 In principle, a distinction between different orientations for helix III (and therefore different confor- mational changes) or multiple orientations that sample a broad range of conformations should be possible through a detailed analysis of existing or new S100 structures. To date, however, this has not been done. In this work, we have determined the water-refined, three-dimensional structure of the calcium-free human S100B dimer at low temperature (108C) to define whether major temperature-induced structural changes occur compared to other S100 structures determined at higher temperatures. We have also used chemical shift in- formation and residual dipolar coupling experiments to determine if more subtle temperature-induced structural changes occur in apo-S100B, and in particular helix III. Using this data, and the structure of human apo-S100B, we have assessed the interhelical interactions and super- positions for the 11 calcium-free S100 protein structures currently available. We find that the apo-S100 proteins exhibit normal Gaussian distributions expected for a sin- gle conformational distribution. These results provide evidence that the structures of apo-S100 proteins reflect their ability to survey a wide range of conformational space rather than adopting a specific conformation. The structure of human apo-S100B is close to the prototypi- cal structure in this population. MATERIAL AND METHODS Sample preparation Uniformly 15 N- and 13 C-labeled recombinant human S100B was expressed in Escherichia coli strain N99, and purified to homogeneity as previously described.49 All NMR experiments utilized 1–2 mM human S100B sam- ples dissolved in 90% H2O/10% D2O containing 50 mM KCl at pH 7.19. DSS was used as an internal standard. A single 15 N/13 C-labeled sample was used for all triple reso- nance NMR experiments. NMR spectroscopy All NMR spectroscopy utilized Varian 600 and 800 MHz spectrometers with pulse field gradient triple reso- nance probes. Data for the backbone sequential assign- ments was collected at 108C and included HNCA,50 HNCACB,51 CBCA(CO)NH,52 HNCO, and 1 H-15 N HSQC experiments. Side chain assignments were made from C(CO)NH, HC(CO)NH,53 HCCH-COSY54 (90% D2O/10% H2O), HBCBCGCDHD,55 and 1 H-13 C HSQC experiments. Interproton distances were measured from 15 N-NOESY (mixing time 150 ms), 13 C-NOESY (mixing time 100 ms, in 90% D2O/10% H2O) and F1-filtered, F3- edited NOESY56 experiments (mixing time 100 ms) were collected at 800 MHz at the Canadian National High Field NMR Centre (NANUC). Slow-exchanging amide protons were identified via a series of 1 H-15 N HSQC experiments taken at time intervals (5 min, 5 h, 10 h, and 16 h) after dissolving the protein in D2O. 1 H-15 N re- sidual dipolar couplings were measured from a series of IPAP-HSQC experiments57–59 using a 0.7 mM sample of 15 N-labeled apo-S100B in 12 mg/mL Pf1 phage with 50 mM KCl at pH 7.19. Spectra were acquired at 35, 25, 15, 10, and 58C. An identical sample lacking phage was used to measure isotropic 1 H-15 N splittings at the same temperatures. All NMR spectra were processed using NMRPipe60 and analyzed using NMRView61 software. Structure calculations Three-dimensional structures of apo-S100B were calcu- lated using the simulated annealing protocol in the pro- gram CNS62 with the use of noncrystallographic symme- try. A total of 2504 nOes, including 1328 intraresidue, 556 sequential, 370 short-range, 160 long-range, and 90 intermolecular distances were used. After the initial Conformational Distribution for Apo-S100 Proteins PROTEINS 29
  • 3. structure calculation was completed to define the second- ary structure of the protein, 124 hydrogen bond distance restraints and 336 dihedral restraints were included in the calculations. Interproton distances for all the proton pairs, except intermolecular nOes, were calibrated using maximum and minimum nOe intensities for known DNHa distances.63 For all intermolecular nOes, an upper distance of 5 A˚ was used. Dihedral angle restraints were determined using the TALOS protocol64 based on chemi- cal shift data from Ha, Ca, Cb, and CO resonances of apo-S100B. Angular restraints were only selected when 9/ 10 dihedral matches fell within an allowed region of the Ramachandran plot. The resulting w and / angles were restricted to two times the error from the TALOS output for the structure calculation with a minimum error of Æ208. Hydrogen bonds were identified from slowly exchanging NH resonances at the 16 h time point from the incubation of apo-S100B in D2O and where tempera- ture coefficients > 24.5 ppb/K could be measured. For each hydrogen bond distance restraints for NHÀÀO (1.8– 2.3 A˚ ) and NÀÀO (2.3–3.3 A˚ ) were used. The final water refinement utilized the RECOORD method65 for CNS. Structures were viewed and analyzed using MOLMOL.66 VADAR67 (University of Alberta) was used to calculate the accessible surface area for all proteins described. All interhelical angles were calculated using the Vector Ge- ometry Mapping method.29,68 In cases where NMR structures were used, calculations were done for multiple structures and the average Æ standard deviation reported. Pairwise comparisons and rmsd calculations for all apo- S100 structures were done using the program Chimera.69 The atomic coordinates for the 20 lowest energy struc- tures of human apo-S100B have been deposited in the RCSB under accession number 2PRU. RESULTS AND DISCUSSION The backbone and side chain resonance assignments for apo-S100B at 108C and 358C were determined by the acquisition and analysis of standard two- and three- dimensional NMR experiments. In total, assignments of 1 H, 13 C, and 15 N resonances for 90 of the 91 residues in apo-S100B were completed. The aromatic side-chain res- onances for all seven phenylalanine, one tyrosine, and four histidine residues in each apo-S100B subunit were accomplished and were critically important for the struc- ture determination of the protein since a large number of intra- and intermolecular nOe correlations were observed to these residues. Temperature dependence of apo-S100B The majority of three-dimensional structures for apo- S100 proteins have been completed using NMR spectros- copy at temperatures above 258C. For example, data used for the structures of apo-S100A6,37,44 S100B40 and S100P70 was acquired at 25–388C. In contrast, the only two X-ray structures of calcium-free S100 proteins, apo- S100A371 and apo-S100A646 had data collected at 21738C, although crystals were grown near ambient temperatures. To determine the influence of temperature on the apo-S100B structure a series of 1 H-15 N HSQC spectra were collected at temperatures ranging between 10–358C. As shown in Figure 1 the majority of the amide correlations shift downfield in both 1 H and 15 N dimen- sions as the temperature is lowered. In general, most of the resonances exhibited some line broadening at the lower temperatures, which is consistent with a longer correlation time due to an increased viscosity/tempera- ture ratio. Further, analysis of the temperature-dependent changes in the cross peak position showed that each res- onance shifted in a linear fashion. The largest change in amide proton chemical shift was $0.2 ppm as the tem- perature was lowered. Similar temperature dependent changes in 1 H chemical shift have been noted for the EF- hand protein troponin-C72 where a large structural change has not been noted. In addition, the changes in chemical shift with decreased temperature for apo-S100B are much less than the 1 H chemical shift changes on cal- cium-binding or interaction with other proteins73,74 where a large structural reorganization of the protein occurs. The temperature coefficients for each amide reso- nance (2DdHN/T) are shown in Figure 2. The data shows Figure 1 1 H-15 N HSQC spectra for human apo-S100B shown as a function of temperature. The spectra are superimposed and plotted using filled contours for 108C and a single contour levels at 25 and 358C. Assignments for most of the isolated correlations are indicated near the 108C spectrum. Arrows are used to show the direction of movement for some of the most affected resonances over the 358C to 108C temperature range. S. Malik et al. 30 PROTEINS
  • 4. the average temperature coefficient is 23.2 ppb/K and most of the residues in apo-S100B are found within one standard deviation (Æ2.3 ppb/K) of this value. Since there are no regions of the sequence that exhibit anoma- lous coefficients this further indicates that large tempera- ture-induced structural changes in apo-S100B do not occur as the temperature is lowered. It has been shown for bovine pancreatic trypsin inhibitor and lysozyme that temperature coefficients more positive than 24.5 ppb/K indicate an amide proton that is hydrogen bonded.75 This appears to hold true for many residues within the helical regions of apo-S100B. However, several amide protons lie in the lower portion of [Fig 2(A)], below the 24.5 ppb/K threshold, indicating poorer hydrogen bond- ing. Many of these are located in unstructured regions, including those that lie either at the extreme C-terminus of the protein (E91) or within calcium-binding sites I (G19, G22, D23, K24, S30) and II (D61, G64, E67, F70). It is interesting to note that two amide resonances that shift greater than one standard deviation from the mean lie at the observed S100 dimerization interface involving helix I (L3, L10). While a loss of intramolecular hydrogen bonding could cause this effect, this result may arise from a minor temperature dependent alteration of the helix I-helix I0 interface. An additional significant change was noted at the N-terminus of helix III (I47, Q50), a region that contains some of the fastest exchanging am- ide protons in apo-S100B.48 To assess whether any re-orientation of helices in apo- S100B occurred between 358C and 108C, we measured HN residual dipolar couplings using two-dimensional IPAP 1 H-15 N HSQC experiments at both temperatures [Fig. 2(B)]. Residual dipolar couplings provide orienta- tional information about the NÀÀH bond vector in a partially oriented sample in a magnetic field. The figure shows the expected clustering of negative couplings for helices I and IV since these helices are nearly coplanar as shown in previously determined S100 structures.40 Over- all, at the two temperatures studied most residual dipolar couplings changed by less than 1 Hz, which is close to the measurement error for the experiments. This is in contrast to much larger changes observed for the ribonu- clease S-peptide, where a significant temperature-induced structural alteration occurs. In apo-S100B, the residual dipolar coupling experiments suggest that little reorienta- tion of any of the helices occurred over the 10–358C temperature range. Alternatively, it is possible that these experiments report only the average position of the heli- ces in apo-S100B and would not be sensitive to helix movement or helices that occupy a range of conforma- tions. To determine this latter case, a more detailed ex- amination of the apo-S100B structure at lower tempera- ture and its comparison with other apo-S100 structures was completed. Description of the structure of human apo-S100B The three-dimensional structure of human apo-S100B was determined at 108C using a total of 2504 nOes, 336 dihedral restraints, and 124 hydrogen bonds. Hydrogen bonds and dihedral restraints were only used in a-helical regions clearly defined from nOe patterns. The amide protons involved in hydrogen bonding were identified from their slow exchange with D2O and measured tem- perature coefficients > 24.5 ppb/K. The 20 lowest energy structures obtained from CNS calculations were used for further refinement using a full nonbonded potential to- gether with explicit solvent molecules. While use of this approach made little difference to the overall structure, Figure 2 A: 1 H temperature coefficients for the backbone amide (HN) resonances of human apo-S100B. The coefficients were determined from the slope of a plot of HN chemical shift vs. temperature between 308 K and 283 K and plotted. The average 1 H temperature coefficient was 23.2 ppb/K (dashed line). Temperature coefficients larger than 24.5 ppb/K (solid line) are consistent with hydrogen bonding. Residues that exhibited temperature coefficients more negative than 24.5 ppb/K, are not expected to be hydrogen bonded and are labeled for clarity. B: HN residual dipolar coupling constants for human apo-S100B measured at 358C (*) and 108C (n). Only residues that exhibited clearly resolved HN correlations were measured. Bars above each figure are shown to represent the positions of the four a-helices, determined in the current work. Conformational Distribution for Apo-S100 Proteins PROTEINS 31
  • 5. improvements were noted in the average nOe and angu- lar restraint violations, and their corresponding energies. In addition, the number of close contacts dropped by nearly 70% in the water-refined approach, a similar result to that noted for other systems.65 The ensemble of the 20 lowest-energy structures of human apo-S100B is shown in Figure 3 and structural statistics are listed in Table I. Structures were chosen based on their low ener- gies and contained no distance violations greater than 0.5 A˚ and no angle violations greater than 58. The family of structures shows two S100B monomers, each with four well-defined helices (Fig. 3: I–IV, I0 –IV0 ). The sym- metric relationship between the monomers is evident from a twofold rotational axis passing through the dimer interface approximately perpendicular to helices I and I0 , and parallel to helices IV and IV0 . The NMR spectra were consistent with this symmetric nature as one resonance was observed for most peaks. Some peaks, such as L3 exhibited multiple peaks (see Fig. 1) because of the pres- ence of formyl- and desformyl methionine forms of the protein that have been previously characterized.76 In all cases, these peaks shifted nearly identically during tem- perature studies (see Fig. 1) indicating the differential N- terminal modification has little affect on dimerization. The rmsd of all helices in the dimer relative to the mean structure was 0.55 Æ 0.07 A˚ for the backbone and 1.03 Æ 0.09 A˚ for all heavy atoms. Each S100B monomer is formed by two helix-loop-he- lix EF-hand calcium-binding motifs joined by a linker region [Fig. 3(A)]. The N-terminal EF-hand consists of helix I (E2-Y17), calcium binding loop I (S18-K28), and helix II (K29-E39). The C-terminal EF-hand comprises helix III (Q50-D60), calcium-binding loop II (D61-D69), and helix IV (F70-T81). Each of the four helices is well defined in all 20 low-energy structures, with the rmsd values 0.33 Æ 0.08 A˚ for helix I, 0.23 Æ 0.07 A˚ for helix Figure 3 Three-dimensional structure of human apo-S100B determined at 108C. A: Ribbon diagram of apo-S100B showing helices I and I0 (blue), II and II0 (yellow), III and III0 (pink), and IV and IV0 (red). B: Superposition (N, Ca, C0 ) of the family of 20 low-energy structures of apo-S100B obtained after water-refinement shown in the same orientation as (A). Table I Structural Statistics of Apo-S100B Restraints for RECOORD refined structure calculation Total NOEs 2504 Intraresidual NOEs 1328 Sequential NOEs 556 Long-range NOEs 530 Intermolecular 90 Dihedral angles /; w (TALOS) 336 Hydrogen bonds 124 Energies (kcal/mol) ETotal 28066.5 Æ 224.8 ENOE 0.81 Æ 0.10 Echid 0.83 Æ 0.50 EL-J 2739.6 Æ 35.2 Structure quality r.m.s.d. from experimental restraints Angular (8) 0.1798 Æ 0.0357 Distance (Š) 0.0173 Æ 0.0015 r.m.s.d. from idealized geometry Bond (Š) 0.0051 Æ 0.0004 Bond angles (8) 1.3380 Æ 0.0637 Improper torsion (8) 1.7632 Æ 0.1050 Ramachandran plot statistics Residues in favored regionsa 94.9% r.m.s.d. to mean structure (Š)b Backbone atoms 0.55 Æ 0.07 Heavy atoms 1.03 Æ 0.09 Bad contacts (avg. per structure) 2.4 a Reflects residues in both most favored and additionally favored regions. b Precision is calculated for residues found in helices I (3–16), II (29–39), III (50–60), and IV (70–80) in apo-S100B. S. Malik et al. 32 PROTEINS
  • 6. II, 0.14 Æ 0.05 A˚ for helix III, and 0.17 Æ 0.03 A˚ for he- lix IV. For the calcium-binding loop regions, long-range nOes were found for residues L27 with C68 and D69, K28 with E67 and C68, and K29 with C68 indicating the two calcium-binding loops were in close proximity. Evidence from nOe data also showed that extensive hydrophobic interactions occurred between the four helices. These interactions include contacts between V13 and F14 (helix I) with L35 (helix II), S30, L32, and I36 (helix II) with V56, M57, and L60 (helix III), V52, V56 (helix III) with F76, M79 and V80 (helix IV), L10 (helix I) with residue F73 (helix IV), and L32, L35, I36 (helix II) with F73 and V80 (helix IV). Analysis of the family of structures indi- cated that 94.9% of all the residues were in the allowed regions of the Ramachandran plot. The dimerization of apo-S100B occurs through an anti- parallel alignment of helices I-I0 and IV-IV0 , and the per- pendicular association of these pairs of helices to form an X-type bundle. Residues that define the dimer interface in human apo-S100B were identified from a 13 C F1-filtered, F3-edited NOESY spectrum using 13 C, 15 N-labelled and unlabelled proteins in a 50:50 ratio in solution. Unambig- uous intermolecular nOes from this data showed interac- tions between A6 (helix I) and A60 , A90 and L100 (helix I0 ), and L3 (helix I) with V130 (helix I0 ). The perpendicular association of helices I and IV0 was evident from unambig- uous intermolecular nOes between L3 and M7 (helix I) with V770 and T810 (helix IV0 ), respectively. The orienta- tion of helices IV and IV0 was determined by nOes between M74 (helix IV) with M740 (helix IV0 ) and, F70 (helix IV) with T820 and F870 (helix IV0 ). Further hydrophobic inter- actions at the dimer interface were observed between the N-terminus of helix I (L3), and the N-terminus of the linker region in the partner monomer (L400 ). The three-dimensional structures of rat40 and bovine43 apo-S100B have been previously determined using NMR spectroscopy. A comparison of these structures reveals a backbone rmsd of 2.00 A˚ (residues 2–84) between the two structures indicating there are some significant differences between these two proteins. This arises mostly from dissimi- larities in the orientation of helix III and calcium-binding loop II. While it is possible sequence variation could con- tribute to this difference, this seems unlikely given that rat and bovine sequences share 95.6% identity (87 of 91 resi- dues). Further, these proteins have identical helix III sequences and a single minor change in calcium-binding loop II (E62 in rat; S62 in bovine). The current structure of human apo-S100B, which shares 97.8% and 96.7% identi- ties with rat and bovine proteins, respectively, allows some of the differences between the rat and bovine structures to be resolved. Using the helices of these apo-S100 proteins for comparison, there is a backbone rmsd of 1.76 A˚ between human and rat apo-S100B that increases to 2.29 A˚ between the human and bovine proteins. These inequities largely result from differences in the positions of helices III and IV where an rmsd of 1.24 A˚ exists between human and rat pro- teins but increases to 2.32 A˚ between human and bovine proteins. Only a single minor, conservative residue change occurs between human and bovine proteins at position 80 (I80 in bovine; V80 in human) in helix IV indicating the sequence difference probably does not account for the larger variance in helix positions between these proteins. Thus, when comparing the a-helices it is clear that human apo- S100B is more similar to the rat S100B structure than the bovine structure. The similarity between the orientation of a-helices in human and rat apo-S100B proteins becomes more apparent when interhelical angles are considered (described later). The human apo-S100B structure (see Fig. 3) indicates that calcium-binding loops I (G19-E32) and II (D61- E72) are less well defined (backbone rmsd 2.27 and 1.18 A˚ , respectively) than the helices. A similar observa- tion has been made for bovine apo-S100B43 although calcium-binding loop I is the better defined of the two loops in that protein. However, in rat apo-S100B,40 the backbone rmsds for calcium-binding loops I (0.22 A˚ ) and II (0.15 A˚ ) are similar to those found in the a-heli- ces suggesting the loops adopt a tight structure with lim- ited flexibility. In human apo-S100B, a network of nOes similar to those observed for helices I-IV was not evident in calcium-binding loops I and II, contributing to its poorer definition. Further, amide exchange experiments show that most of the amide protons within calcium- binding loops I and II have protection factors up to six orders of magnitude lower than amide protons in the a- helices.48 This indicates that a poorer series of hydrogen bonds exists in the calcium-binding loops of human apo- S100B and that most of the amide protons are exposed to solvent. In addition, 15 N relaxation experiments77 have shown that the order parameters (S2 ) for residues in calcium-binding loops I and II average 0.80, perhaps indicating a greater degree of flexibility within the loops than the helices (S2 5 0.84–0.87). It is interesting that 15 N relaxation experiments with another EF-hand cal- cium-binding protein troponin-C,78,79 have indicated calcium-binding sites I and II in the N-terminal domain have lower order parameters also, characteristic of greater flexibility. Thus, the poor definition of the calcium-bind- ing loops in human apo-S100B is most consistent with flexibility of the backbone within these regions. This interpretation is in agreement with crystallographic stud- ies of apo-S100A371 and apo-S100A646 where thermal factors about two-times those found in the helices have been observed for site I in apo-S100A6 and site II in apo-S100A3, corresponding to backbone atomic displace- ments near 0.6 A˚ . The apo-S100 proteins show variations in interhelical angles All S100 proteins, with the exception of S100A10, undergo a calcium-induced conformational change that Conformational Distribution for Apo-S100 Proteins PROTEINS 33
  • 7. allow for their interaction with a variety of target pro- teins. In general, this change involves a reorientation of helix III and repacking of helix II that results in the ex- posure of a broad, hydrophobic surface. To assess the details of this conformational change and to determine whether specific S100 proteins adapt to calcium binding in different manners, knowledge of the structures of many S100 protein family members is required in the apo- and calcium-bound states. We have used the struc- ture of human apo-S100B to assess the detailed structure of all S100 proteins in the calcium-free state to under- stand the first part of the calcium-binding response. The structure of human apo-S100B determined at 108C was compared with NMR and X-ray crystallo- graphic structures of S100A1, S100A3, S100A4, S100A6, S100A11, S100A13, bovine and rat S100B, and S100P using a variety of criteria, including interhelical distances and angles, helix rotation and buried surface area. The helix–helix relationship in this broad range of S100 struc- tures was examined using the vector geometry mapping (VGM) method, which provides important information about the tip angle (y) between helices, the helix projec- tion angle (/) and the helix role (x), and gives the best overall picture of the spatial arrangement of helices.29,68 In our assessment, first shown for apo-calmodulin, we chose helices adjacent to the calcium-binding loops such that the incoming helix ended at the hydrophobic residue immediately preceding the calcium-binding loop. In apo- S100B this corresponded to Y17 and L60 in helices I and III. The exiting helix started two residues before the bidentate glutamate-coordinating residue that terminates the calcium-binding loop (E31 and E72 in helices II and IV of human S100B). In all cases, a careful examination was made of the helix selections by shifting the helix des- ignation by one or two residues to insure the y and / were not grossly affected. It was observed that some heli- ces, especially helices I and III, were very sensitive to this selection due to the presence of some helix twist or bend. The interhelical angles y and / for human apo-S100B and 10 other apo-S100 structures are listed in Table II and shown graphically in Figure 4. The y and / angles between helices I and II for the pseudo EF-hand (EF1) in human apo-S100B are 72 Æ 48 and 98 Æ 88, respectively. These are in excellent agreement with the average angles (63 Æ 108, 99 Æ 78) obtained when all other apo-S100 structures are considered, indicating that helices I and II adopt near identical orientations amongst all available apo-S100 protein structures. This is shown clearly in [Fig. 4(A)], where there is a clustering of the exiting he- lix II for each structure. In addition, the pairwise super- positions for helices I and II in these structures (55 com- binations) was binned and fit with a Gaussian distribu- tion having a midpoint of 1.46 Æ 0.56 A˚ [Fig. 5(A)]. This data passed multiple tests for a normal distribution. The excellent agreement of the midpoint for the Gaussian fit, the median (1.46) and positive normal distribution tests was strong evidence for a single structural popula- tion. This would indicate that the distribution of angles for helices I and II in the apo-S100 structures is most consistent with a single conformational family. Further, the range of angles for EF1 in the apo-S100 proteins (Ta- ble II) is considerably tighter than that observed for heli- ces I and II in all EF-hand protein structures (1158).80 A comparison of helices I and II in EF1 from the S100 family to apo-calmodulin shows that calmodulin has a significantly smaller tip angle (y) of 46 Æ 28, and larger horizontal angle (/) of 123 Æ 38, both of which lie out- side the ranges observed for any single S100 protein. As shown in Figure 4, the position of the C-terminus of he- lix II is closer to the N-terminus of helix I for apo-cal- modulin than for any of the S100 proteins, resulting in y and / angles for apo-calmodulin that differ by about 2178 and 248, respectively, when compared with the av- erage apo-S100 structure. Human apo-S100B and the rest of the S100 proteins thus have a more open conforma- tion for EF1 than apo-calmodulin, a result reaffirmed when considering interhelical distances between helices I and II of EF1. The separation of these helices in the S100 proteins is approximately 3.6 A˚ greater than those in apo-calmodulin. The tip and horizontal angles for EF2 of human apo- S100B are y 5 25 Æ 28 and / 5 2146 Æ 88, respectively (Table II). These are well within the range observed for all apo-S100 proteins (y 5 25 Æ 88, / 5 2142 Æ 468). In general the exiting helix IV in the apo-S100B proteins is angled in an opposite direction compared with apo- calmodulin (see Fig. 4), which has a positive / angle most similar to S100P. Unlike EF1, the larger deviations in the y and / angles for EF2 indicate that potential dif- ferences might exist between the structures due to sequence differences, helix–helix interactions or both. Alternatively, the differences in the structures could reflect a lack of definition in the NMR structures due to limited numbers of nOes between helices III and IV. This does not appear to be the case since most of the apo- S100 structures have tight ranges of y and / angles ( 208) indicating the nOe distance information used in these structures was sufficient. In addition, variation is noted for the two x-ray structures that are available (S100A3 and S100A6) that have / angles that differ by >208, although this difference is clearly smaller than that exhibited by the apo-S100 structures determined by NMR spectroscopy. An examination of the y and / angles for the apo- S100 structures shows that, with the exception of S100P, all the apo-S100 structures are found within 88 of the av- erage y angle with approximately equal numbers on ei- ther side of the average. This indicates that the opening angles between helices III and IV are similar as noted by the mostly parallel arrangement of the helices for the apo-S100 proteins [Fig. 4(D)]. On the other hand, the S. Malik et al. 34 PROTEINS
  • 8. TableII EF-HandAnglesofApo-S100Proteinsa Protein EF1b EF2 N-terminalcoordinate ofsecondhelix y (degree) / (degree) x (degree) N-terminalcoordinateof secondhelix y (degree) / (degree) x (degree) hS100Bc 11.8Æ0.7,23.3Æ1.7,21.1Æ0.772Æ498Æ8129Æ88.4Æ1.1,11.5Æ0.7,21.8Æ0.525Æ22146Æ883Æ8 bS100Bd 10.3Æ1.7,1.6Æ1.5,25.4Æ2.153Æ2105Æ493Æ41.4Æ1.3,10.8Æ1.4,28.2Æ1.310Æ42124Æ1045Æ3 rS100Be 12.3Æ0.3,24.0Æ1.2,25.2Æ0.758Æ193Æ3129Æ47.8Æ0.6,10.4Æ0.5,24.6Æ0.515Æ12127Æ595Æ4 S100A1f 13.1Æ1.3,23.3Æ2.4,24.7Æ1.368Æ391Æ5166Æ63.3Æ1.9,14.3Æ0.5,20.3Æ0.333Æ1284Æ473Æ8 S100A3g 11.2,20.6,24.25510212911.7,10.9,23.0262116107 S100A4h 12.4Æ1.2,22.2Æ1.4,23.2Æ0.768Æ2105Æ3114Æ48.1Æ2.8,11.9Æ2.1,23.7Æ1.419Æ22151Æ770Æ10 # S100A6i 10.1Æ0.7,0.4Æ0.3,24.8Æ0.556Æ1103Æ3113Æ310.1Æ0.8,11.0Æ0.3,22.0Æ0.434Æ22161Æ395Æ3 S100A6j 10.361,20.181,24.8095610512111.4,11.4,23.5242138111 S100A11k 9.9Æ0.4,21.3Æ0.7,24.3Æ0.363Æ2101Æ1123Æ88.7Æ2.9,6.9Æ1.3,22.2Æ1.230Æ4163Æ1073Æ20 S100Pl 10.8Æ0.7,21.0Æ0.7,25.9Æ1.154Æ3105Æ4104Æ29.0Æ0.2,1.49Æ0.7-4.7Æ0.570Æ3121Æ386Æ4 S10013m 13.5Æ1.5,28.3Æ2.5,20.3Æ1.284Æ684Æ9180Æ92.2Æ1.4,14.3Æ0.5,20.5Æ1.030Æ3279Æ757Æ6 AverageS10063Æ1099Æ7128Æ2529Æ15(25Æ8)2142Æ4681Æ20 CaM10.6Æ0.6,20.7Æ0.7,23.2Æ0.346Æ2123Æ3104Æ510.9Æ0.9,5.6Æ1.6,21.8Æ1.048Æ4139Æ489Æ4 a AngleswerecalculatedusingVectorGeometryMapping(VGM)methodusingapo-calmodulin(PDBentry1DMO)asthereference.29,67 b HelicesdefinedforEF1arehelixIandIIandforEF2ashelixIIIandIVfor:S100B-L10–Y17,K29–E39,V53–L60,F70–V80;S100A1-L11–H18,K30–E40V54–L61,F71–L81;S100A3-I12–Y19,Q31–E41Y55–L62,F72–L82; S100A4-M12–Y19,K31–E41F55–L62,F72–I82;S100A6-L12–Y19,K31–L41,I53-L60(# E52–D59),F70–L80;S100A11-L13–Y20,K32–E42,L56–L63,F73–L83;S100P-I11–Y18,K30–E40V54–L61,F71–I81;S100A13-V16-F23, V35-Q45,L56-L63,F73-L83. PDBentriesare c 2PRU d 1CFP e 1B4C f 1K2H g 1KSO h 1M31 i 2CNP j 1K9P k 1NSH l 1OZO m YUS. Conformational Distribution for Apo-S100 Proteins PROTEINS 35
  • 9. differences in the positions of helix IV with respect to helix III mostly arise from the large variation in / angle. For example, some structures (S100A13, S100A1) have / angles about 608 more positive and others (S100A11, S100P) have / angles about 608 more negative than found in human apo-S100B. The VGM plot in Figure 4 fixes helix III along the z-axis and gives the impression that the N-termini of helix IV, closest to the calcium- binding loop of EF2 takes on a range of positions with respect to helix III. Superposition of the apo-S100 pro- teins, however, shows that it is the C-terminus of helix III that occupies a variety of locations. Despite the similarity in y angles observed for the pro- teins it is tempting to suggest that the range of / angles observed for the EF2 calcium-binding site might arise from different structural subfamilies in the apo-S100 pro- teins. In turn, the different interhelical relationships between helices III and IV might be important for the extent that these helices change conformation upon cal- cium binding. Several factors indicate this is not the case. For example, statistical analysis of the pairwise superposi- tions of helices III and IV in EF2 for human apo-S100B and all other apo-S100 proteins shows a Gaussian distri- bution [Fig. 5(B)] that is satisfied at the 95% confidence level and passes several normality tests (with the excep- tion of S100P). Further, the median of this superposition (1.85 A˚ ) falls near the midpoint for a Gaussian distribu- tion of the helices (1.88 Æ 0.73 A˚ ). This result is similar to that observed for helices I, II and IV (1.72 Æ 0.48 A˚ ) or all four helices (1.84 Æ 0.63 A˚ ) shown in Figure 5. Using the pairwise rmsd comparisons (see Fig. 5) as a guide in combination with the VGM analysis (see Fig. 4), it is apparent the large structural differences arise due to the displacement of the N-termini of helix IV with respect to the C-termini of helix III. For example, human apo-S100B and S100A1 have a relatively small rmsd Figure 4 Helix orientation for human apo-S100B and other S100 proteins obtained using Vector Geometry Mapping.29,68 The figure shows helices I and II (A, B) from calcium- binding site EF1 and helices III and IV (C, D) from calcium-binding site EF2. In both cases the incoming helix of each structure is aligned along the z-axis with respect to the N-terminal helix of EF1 in calmodulin. The exiting helices (II in EF1 and IV in EF2) are shown for the calcium-free proteins calmodulin (grey), human S100B (red), bovine S100B (brown), rat S100B (maroon), S100A1 (cyan), S100A3 (green), S100A4 (magenta), human S100A6 (blue), rabbit S100A6 (light blue), S100A11 (dark purple), S100A13 (orange), and S100P (yellow). The plots show the helices from calcium-binding site EF1 as viewed down the 1z axis (A, C) and rotated $908 about y and $458 about z (B, D). The directions of the helices are denoted by N?C either along the z-axis or the exiting helix (B, D). S. Malik et al. 36 PROTEINS
  • 10. when helices III and IV are superimposed (1.30 A˚ ) yet the N-termini of helix IV are separated by about 6 A˚ . This is evident in [Fig. 4(C,D)] where the C-termini of helix IV from human apo-S100B (red) and apo-S100A1 (cyan) are close together, but their N-termini are more divergent. The comparison of the interhelical y and / angles of the apo-S100B structures shows that human apo-S100B lies very close to the midpoint of tip and horizontal angles for both EF1 and EF2 calcium-binding sites. Anal- ysis of the angular ranges and pairwise backbone rmsd indicates that all apo-S100 structures to date, with the exception of S100P (y angle for EF2), fall within a single Gaussian distribution although the ranges of the horizon- tal angle (/) for EF2 is nearly sevenfold larger for EF2 than for EF1. Helix III shows the broadest range of accessible surface areas in the S100 proteins The VGM results indicate there is a significant range of conformations that helices III and IV adopt with respect to each other within the S100 family. In order to pinpoint a rationale for this, the accessible surface area for human apo-S100B, presented here, and other cal- cium-free S100 proteins was analyzed. The fraction acces- sible surface area for human apo-S100B reveals a large number of residues in helix I (L3, A6, M7, A9, L10, I11, V13, and F14) and helix IV (F70, F73, M74, F76, V77, A78, V80, and T81) that have >80% of their side chains buried (see Fig. 6). Most of these residues form the hydrophobic dimer interface in apo-S100B derived from the near perpendicular arrangement of helices I and I0 Figure 5 Comparison of apo-S100 structures showing the distribution of structures as a function of backbone rmsd between helices. The figure shows (A) helices I and II, (B) helices III and IV, (C) helices I, II, and IV, and (D) sum of all helices for human apo-S100B and the other ten apo-S100 structures described in Table II. The helices used were L3-Y17 (helix I), K29-E39 (helix II), Q50-L60 (helix III), and F70-V80 (helix IV) in human apo-S100B and the corresponding regions in other S100 proteins based on alignment using T-Coffee.81 For NMR structures, the most representative structure as listed in the PDB file was used. In each case the rmsd values between all possible pairs of structures (55 comparisons) were tabulated and binned (0.2 A˚ bins). The number of occurrences was plotted against the bin center. Each graph shows the best-fit Gaussian curve to the binned rmsd data centered at (A) 1.46 Æ 0.56 A˚, (B) 1.88 Æ 0.73 A˚, (C) 1.72 Æ 0.48 A˚, and (D) 1.84 Æ 0.63 A˚. The structures for apo- S100A13 and apo-S100P were removed from datasets (A) and (B–D), respectively based on ANOVA statistics. Conformational Distribution for Apo-S100 Proteins PROTEINS 37
  • 11. with IV and IV’. For example, L3 in helix I interacts with L10 and V13 in helix I0 . In helix IV, residue F73 interacts with L3 of helix I, and L10 and F14 of helix I0 . Compari- son with other apo-S100 structures reveals this pattern of buried hydrophobic residues is nearly perfectly preserved [Fig. 6(A,D)]. An examination of the S100 sequences along with this analysis indicates the dimerization motif for these S100 proteins is LXX[A/C/S][M/L/I/V]XX[M/L/ I/V][I/V]X[V/I/T]F (residues 3-14; helix I) and FXE[F/ Y][V/I/L/M]X[L/F][V/L/I][A/G/S]X[V/L/I][T/A] (residues 70–81; helix IV). The excellent agreement between the buried surface area and sequence conservation of the apo- S100 structures is in accord with 15 N relaxation77 and am- ide proton exchange experiments48 for apo-S100B. In par- Figure 6 Comparison of side chain fractional accessible surface area (FASA) for residues in human apo-S100B and other apo-S100 structures. Residues in (A) helix I, (B) helix II, (C) helix III and (IV) helix IV are plotted as a function of fraction accessible surface area by each side chain. The numbering scheme used is that of human apo-S100B and other proteins were aligned based on sequence using T-Coffee.86 The proteins shown are human S100B (n), bovine S100B (~), rat S100B (!), S100A1 (^), S100A3 (l), S100A4 (h), human S100A6 (!), rabbit S100A6 (~), S100A11 (^), S100A13 (3) and S100P (*). All fractional accessible surface areas were calculated from available PDB coordinates using the program VADAR.67 For NMR structures, the most representative structure as listed in the PDB file was used. S. Malik et al. 38 PROTEINS
  • 12. ticular, helices I and IV had protection factors that were similar to free energies for the unfolding of apo-S100B indicating the dimer interface, maintained by the helix I and IV motifs is a major contributor towards the stability of all S100 proteins. Extending this analysis to helix II shows the hydropho- bic portions of residues K29, L32, L35, and I36 are all buried >80% in human apo-S100B. As with helices I and IV these positions in helix II are highly conserved in the S100 sequences. The large variation in accessible sur- face area at the N-terminus of helix II is also accompa- nied by decreased protection from amide exchange.48 In helix III, the shortest of the helices in apo-S100B, a large variation in the accessible surface area exists between structures. Only two residues (V56, L60) show a consist- ent pattern of burial between human apo-S100B and the other S100 proteins and these are both accompanied by high conservation at these positions. Other residues are less well conserved and show poorer patterns of accessi- ble surface area. For example, V52 in the apo-S100B structures is nearly completely buried while residues in S100A6 (E52) and S100A3 (D54) are more exposed con- sistent with their significant differences in side chain po- larity. Notably, position 57 in helix III, where methionine is highly conserved, shows an inconsistent pattern. In human apo-S100B it shows about 30% accessible surface area, midway between the observed range for this residue (2%–62%), despite the conserved nature of the methio- nine residue in all sequences but S100A13 and S100P. Further differences are seen at the second position in the calcium-binding loop occupied by N62 in human apo- S100B, where it is exposed but nearly completely buried in S100A11 where a leucine residue occupies this posi- tion. The large variations in buried surface area through- out helix III and especially for conserved residues (V53, M57) are most consistent with this helix being loosely packed. This conclusion is borne out by amide hydrogen exchange rates for residues in helix III that are 2–3 orders of magnitude faster than helices I, II, and IV.48 In addi- tion, although amide exchange rates have not been meas- ured for rat or bovine apo-S100B, 15 N relaxation experi- ments indicate helix III in rat apo-S100B has the lowest order parameters of the four helices.77 A similar obser- vation has been made in the N-terminal domain of tro- ponin-C78,79 where helix C, one of the helices that undergoes a conformational change upon calcium bind- ing, has the lowest order parameters of the helices in that domain. Function relevance for the S100 proteins An analysis of the tip and horizontal angles for the EF1 and EF2 calcium-binding sites in human apo-S100B and several other apo-S100 proteins reveals a range of orientations are adopted by EF2 with a mean near that determined for human apo-S100B. It was also shown that helix III is clearly the most loosely packed of the four helices in all structures of apo-S100 proteins. This data indicates that helix III is likely a flexible helix able to sample a range of orientations with respect to helix IV. Is this important for the function of the S100 pro- teins? As shown in Figure 7, the broad range of tip and horizontal angles displayed in the apo-S100 structures is not exhibited in the calcium-bound forms. As with the apo-proteins the structure of the calcium-bound form of human S100B has tip and horizontal angles for EF2 (818, 100.18) that are representative of the average angles of 67.5 Æ 9.5 and 100.3 Æ 8.88 for the calcium-bound structures. For human S100B, this indicates that the pro- tein undergoes changes of about 558 and 1148 in the tip and horizontal angles respectively in response to calcium binding. This opening of the horizontal angle is 2.5–3.0 times larger than observed for calmodulin82 or troponin-C.83,84 This structural change is responsible for the shallow binding surface noted in the calcium- bound forms of the S100 proteins compared to a nar- rower cleft found in calmodulin and troponin-C.85,86 It is interesting that the calcium-bound S100 structures compared (4 NMR, 6 X-ray structures) display a much narrower range of horizontal angles (100.3 Æ 8.88) clearly indicating a more homogeneous group of structures compared to their apo forms (Table II). This would indi- cate that calcium binding might act to restrict the freedom of helix III, presumably through an ordering of the EF2 cal- cium-binding site. For S100B, amide exchange experiments Figure 7 Helix reorientation for the EF2 calcium-binding site in S100 proteins. The relative distribution for each S100 protein is plotted versus the horizontal angle (/) calculated using Vector Geometry Mapping.29,68 NMR structures are represented as a Gaussian curve centered at the mean and having a width at half-height corresponding to Æ the standard deviation. X-ray structures are shown as vertical sticks. The structures shown are; human S100B (red), bovine S100B (orange), rat S100B (maroon), S100A1 (green), S100A3 (ochre), S100A4 (blue), human S100A6 (sky blue), rabbit S100A6 (dark green), S100A11 (magenta), S100A12 (pink) S100A13 (rose), and S100P (black). Conformational Distribution for Apo-S100 Proteins PROTEINS 39
  • 13. have shown that the amide rates of exchange in this cal- cium-binding site are slowed by 1–2 orders of magni- tude. However, amide exchange rates for helix III remain the fastest of the helices in S100B indicating this helix is still exposed.48 Still, Figure 7 provides evidence that cal- cium binding to the S100 proteins enables EF2 to move from a broad range of conformations to a tighter more compressed distribution. CONCLUSIONS Based on amide chemical shift changes and residual dipolar changes the three-dimensional structure of human apo-S100B does not appear to undergo signifi- cant structural changes between 35 and 108C. A compari- son of the three dimensional structure of human apo- S100B with those of rat and bovine apo-S100B indicates the human form has similar helix orientations and preci- sion as the rat form of the protein but has calcium-bind- ing loops more similar to those obtained for the bovine structure. Using human apo-S100B as a template the interhelical angles for EF-hands EF1 and EF2 were com- pared with other apo-S100 protein structures. The arrangements of EF1 (y 5 72 Æ 4, / 5 98 Æ 8) and EF2 (y 5 25 Æ 2, / 5 2146 Æ 8) were found to be near the mean for all apo-S100 protein structures deter- mined to date. Although a broad distribution of the EF2 helical angles exist for the apo-S100 proteins, we have shown that these structures assume a normal Gaussian distribution about this conformation indicative of a sin- gle structural population. Calcium binding alters the con- formation of the EF2 helices (III, IV) but more signifi- cantly leads to a tighter less variable arrangement of the helices. ACKNOWLEDGMENTS The authors would like to thanks Kathryn Barber (UWO) for her technical support. We are grateful to Lewis Kay (University of Toronto) for providing pulse sequences, Frank Delaglio for NMRPipe and DYNAMO, and Bruce Johnson for NMRView. We would like to thank the Canadian National High Field NMR Centre (NANUC) for their assistance and use of the facilities. Operation of NANUC is funded by the Canadian Insti- tutes of Health Research, the Natural Science and Engi- neering Research Council of Canada and the University of Alberta. REFERENCES 1. Santamaria-Kisiel L, Rintala-Dempsey AC, Shaw GS. Calcium-de- pendent and -independent interactions of the S100 protein family. Biochem J 2006;396:201–214. 2. Schafer BW, Wicki R, Englekamp D, Mattei MG, Heizmann CW. Isolation of a YAC clone covering a cluster of nine S100 genes on human chromosome 1q21: rationale for a new nomenclature of the S100 calcium-binding protein family. Genomics 1995;25:638– 643. 3. Vogl T, Ludwig S, Goebeler M, Strey A, Thorey IS, Reichelt R, Foell D, Gerke V, Manitz MP, Nacken W, Werner S, Sorg C, Roth J. MRP8 and MRP14 control microtubule reorganization during transendothelial migration of phagocytes. Blood 2004;104:4260– 4268. 4. Frizzo JK, Tramontina AC, Tramontina F, Gottfried C, Leal RB, Donato R, Goncalves CA. Involvement of the S100B in cAMP- induced cytoskeleton remodeling in astrocytes: a study using TRTK-12 in digitonin-permeabilized cells. Cell Mol Neurobiol 2004;24:833–840. 5. Bianchi R, Garbuglia M, Verzini M, Giambanco I, Ivanenkov VV, Dimlich RV, Jamieson GA, Jr, Donato R. S-100 (alpha and beta) binding peptide (TRTK-12) blocks S-100/GFAP interaction: identi- fication of a putative S-100 target epitope within the head domain of GFAP. Biochim Biophys Acta 1996;1313:258–267. 6. Davey GE, Murmann P, Hoechli M, Tanaka T, Heizmann CW. Cal- cium-dependent translocation of S100A11 requires tubulin fila- ments. Biochim Biophys Acta 2000;1498:220–232. 7. Garbuglia M, Verzini M, Rustandi RR, Osterloh D, Weber DJ, Gerke V, Donato R. Role of the C-terminal extension in the interac- tion of S100A1 with GFAP, tubulin, the S100A1- and S100B-inhibi- tory peptide. TRTK-12, and a peptide derived from p53, and the S100A1 inhibitory effect on GFAP polymerization. Biochem Bio- phys Res Comm 1999;254:36–41. 8. Deora AB, Kreitzer G, Jacovina AT, Hajjar KA. An annexin 2 phos- phorylation switch mediates p11-dependent translocation of annexin 2 to the cell surface. J Biol Chem 2004;279:43411–43418. 9. Bichsel SJ, Tamaskovic R, Stegert MR, Hemmings BA. Mechanism of activation of NDR (nuclear Dbf2-related) protein kinase by the hMOB1 protein. J Biol Chem 2004;279:35228–35235. 10. Wilder PT, Lin J, Bair CL, Charpentier TH, Yang D, Liriano M, Var- ney KM, Lee A, Oppenheim AB, Adhya S, Carrier F, Weber DJ. Rec- ognition of the tumor suppressor protein p53 and other protein targets by the calcium-binding protein S100B. Biochim Biophys Acta 2006;1763:1284–1297. 11. Sheu F-S, Huang FL, Huang K-P. Differential responses of protein kinase C substrates (MARCKS, neuromodulin and neurogranin) phosphorylation to calmodulin and S100. Arch Biochem Biophys 1995;316:335–342. 12. Baudier J, Bergeret E, Bertacchi N, Weintraub H, Gagnon J, Garin J. Interactions of myogenic bHLH transcription factors with cal- cium-binding calmodulin and S100a (alpha alpha) proteins. Bio- chemistry 1995;34:7834–7846. 13. Baudier J, Bronner C, Kligman D, Cole RD. Protein kinase C sub- strates from bovine brain. J Biol Chem 1989;264:1824–1828. 14. Hatakeyama T, Okada M, Shimamoto S, Kubota Y, Kobayashi R. Identification of intracellular target proteins of the calcium-signal- ing protein S100A12. Eur J Biochem 2004;271:3765–3775. 15. Zimmer DB, Van Eldik LJ. Identification of a molecular target for the calcium-modulated protein S100. Fructose-1,6-bisphosphate al- dolase. J Biol Chem 1986;261:11424–11428. 16. Landar A, Caddell G, Chessher J, Zimmer DB. Identification of an S100A1/S100B target protein: phosphoglucomutase. Cell Calcium 1996;20:279–285. 17. Lewit-Bentley A, Rety S, Sopkova-de OliveiraSantos J, Gerke V. S100-annexin complexes: some insights from structural studies. Cell Biol Int 2000;24:799–802. 18. Seemann J, Weber K, Gerke V. Annexin I targets S100C to early endosomes. FEBS Lett 1997;413:185–190. 19. Gerke V. Tyrosine protein kinase substrate p36: a member of the annexin family of Ca21 /phospholipid-binding proteins. Cell Motil Cytoskeleton 1989;14:449–454. 20. Arcuri C, Giambanco I, Bianchi R, Donato R. Annexin V, annexin VI, S100A1 and S100B in developing and adult avian skeletal muscles. Neuroscience 2002;109:371–388. S. Malik et al. 40 PROTEINS
  • 14. 21. Chang N, Sutherland C, Hesse E, Winkfein R, Wiehler WB, Pho M, Veillette C, Li S, Wilson DP, Kiss E, Walsh MP. Identification of a novel interaction between the Ca(21)-binding protein S100A11 and the Ca(21)- and phospholipid-binding protein annexin A6. Am J Physiol Cell Physiol 2007;292:C1417–C1430. 22. Nowotny M, Spiechowicz M, Jastrzebska B, Filipek A, Kitagawa K, Kuznicki J. Calcium-regulated interaction of Sgt1 with S100A6 (cal- cyclin) and other S100 proteins. J Biol Chem 2003;278:26923– 26928. 23. Filipek A, Jastrzebska B, Nowotny M, Kuznicki J. CacyBP/SIP, a cal- cyclin and Siah-1-interacting protein, binds EF-hand proteins of the S100 family. J Biol Chem 2002;277:28848–28852. 24. Fujii T, Hiromori T, Hamamoto M, Suzuki T. Interaction of chicken gizzard smooth muscle calponin with brain microtubules. J Biochem (Tokyo) 1997;122:344–351. 25. Baudier J, Delphin C, Grunwald D, Khochbin S, Lawrence JJ. Char- acterization of the tumor suppressor protein p53 as a protein kinase C substrate and a S100b-binding protein. Proc Natl Acad Sci U S A 1992;89:11627–11631. 26. Lin J, Blake M, Tang C, Zimmer D, Rustandi RR, Weber DJ, Carrier F. Inhibition of p53 transcriptional activity by the S100B calcium- binding protein. J Biol Chem 2001;276:35037–35041. 27. Scotto C, Deloulme JC, Rousseau D, Chambaz E, Baudier J. Cal- cium and S100B regulation of p53-dependent cell growth arrest and apoptosis. Mol Cell Biol 1998;18:4272–4281. 28. Yu WH, Fraser PE. S100B interaction with Tau is promoted by zinc and inhibited by hyperphosphorylation in Alzheimer’s disease. J Neurosci 2001;21:2240–2246. 29. Yap KL, Ames JB, Swindells MB, Ikura M. Diversity of conforma- tional states and changes within the EF-hand protein superfamily. Proteins 1999;37:499–507. 30. Ikura M. Calcium binding and conformational response in EF-hand proteins. Trends Biochem 1996;21:14–17. 31. Marshak DR, Pesce SA, Stanley LC, Griffin WST. Increased S100b neurotrophic activity in Alzheimer’s disease temporal lobe. Neuro- biol Aging 1991;13:1–7. 32. Boom A, Pochet R, Authelet M, Pradier L, Borghgraef P, Van Leuven F, Heizmann CW, Brion JP. Astrocytic calcium/zinc binding protein S100A6 over expression in Alzheimer’s disease and in PS1/ APP transgenic mice models. Biochim Biophys Acta 2004;1742:161– 168. 33. Yui S, Nakatani Y, Mikami M. Calprotectin (S100A8/S100A9), an inflammatory protein complex from neutrophils with a broad apo- ptosis-inducing activity. Biol Pharm Bull 2003;26:753–760. 34. Odink K, Cerletti N, Bruggen J, Clerc RG, Tarcsay L, Zwadlo G, Gerhards G, Schlegel R, Sorg C. Two calcium-binding proteins in infiltrate macrophages of rheumatoid arthritis. Nature 1987;330:80– 82. 35. Deloulme JC, Gentil BJ, Baudier J. Monitoring of S100 homodime- rization and heterodimeric interactions by the yeast two-hybrid sys- tem. Microsc Res Tech 2003;60:560–568. 36. Deloulme JC, Assard N, Mbele GO, Mangin C, Kuwano R, Baudier J. S100A6 and S100A11 are specific targets of the calcium- and zinc-binding S100B protein in vivo. J Biol Chem 2000;275:35302– 35310. 37. Potts BCM, Smith J, Akke M, Macke TJ, Okazaki K, Hidaka H, Case DA, Chazin WJ. The structure of calcyclin reveals a novel homodimeric fold for S100 Ca21 -binding proteins. Nature Struct Biol 1995;2:790–796. 38. Rety S, Sopkova J, Renouard M, Osterloh D, Gerke V, Tabaries S, Russo-Marie R, Lewit-Bentley A. The crystal structure of a complex of p11 with the annexin II N-terminal peptide. Nature Struct Biol 1999;6:89–95. 39. Rety S, Osterloh D, Arie JP, Tabaries S, Seeman J, Russo-Marie F, Gerke V, Lewit-Bentley A. Structural basis of the Ca21 -dependent association between S100C (S100A11) and its target, the N-terminal part of annexin I. Structure 2000;8:175–184. 40. Drohat AC, Tjandra N, Baldisseri DM, Weber DJ. The use of dipo- lar couplings for determining the solution structure of rat apo- S100B(bb). Protein Sci 1999;8:800–809. 41. Drohat AC, Baldisseri DM, Rustandi RR, Weber DJ. Solution structure of calcium-bound rat S100B(betabeta) as determined by nuclear mag- netic resonance spectroscopy. Biochemistry 1998;37:2729–2740. 42. Smith SP, Shaw GS. A novel calcium-sensitive switch revealed by the structure of human S100B in the calcium-bound form. Struc- ture 1998;6:211–222. 43. Kilby PM, Van Eldik LJ, Roberts GCK. The solution structure of the bovine S100b protein dimer in the calcium-free state. Structure 1996;4:1041–1052. 44. Maler L, Potts BC, Chazin WJ. High resolution solution structure of apo calcyclin and structural variations in the S100 family of cal- cium-binding proteins. J Biomol NMR 1999;13:233–247. 45. Sastry M, Ketchem RR, Crescenzi O, Weber C, Lubienski MJ, Hidaka H, Chazin WJ. The three-dimensional structure of Ca21 - bound calcyclin: implications for Ca21 -signal transduction by S100 proteins. Structure 1998;6:223–231. 46. Otterbein LR, Kordowska J, Witte-Hoffmann C, Wang C-L, Domi- nguez R. Crystal structures of S100A6 in the Ca21 -free and Ca21 - bound states: the calcium sensor mechanism of S100 proteins revealed at atomic resolution. Structure 2002;10:557–567. 47. Dempsey AC, Walsh MP, Shaw GS. Unmasking the annexin I inter- action from the structure of Apo-S100A11. Structure (Camb) 2003;11:887–897. 48. Marlatt NM, Shaw GS. Amide exchange shows calcium-induced conformational changes are transmitted to the dimer interface of S100B. Biochemistry 2007;46:7478–7487. 49. Smith SP, Barber KR, Dunn SD, Shaw GS. Structural influence of cation binding to recombinant human brain S100b: evidence for calcium-induced exposure of a hydrophobic surface. Biochemistry 1996;35:8805–8814. 50. Bax A, Ikura M. An efficient 3D NMR technique for correlating the proton and 15N backbone amide resonances with the a-carbon of the preceding residue in uniformly 15N/13C enriched proteins. J Biomol NMR 1991;1:99–104. 51. Wittekind M, Mueller L. HNCACB, A high sensitivity 3D NMR experiment to correlate amide proton and nitrogen resonances with the a-carbon and b-carbon resonances in proteins. J Magn Reson Series B 1993;101:171–180. 52. Grzesiek S, Bax A. An efficient experiment for sequential backbone assignment of medium-sized isotopically enriched proteins. J Magn Reson 1992;99:201–207. 53. Grzesiek S, Anglister J, Bax A. Correlation of backbone amide and aliphatic side-chain resonances in 13 C/15 N-enriched proteins by iso- tropic mixing of 13 C magnetization. J Magn Reson Series B 1993;101:114–119. 54. Kay LE, Xu G, Singer AU, Muhandiram DR, Forman-Kay JD. A gradient-enhanced HCCH-TOCSY experiment for recording side- chain 1 H and 13 C correlations in H2O samples of proteins. J Magn Reson 1993;101:333–337. 55. Yamazaki T, Forman-Kay JD, Kay LE. Two-dimensional NMR experiments for correlating 13 Cb and 1 Hd/e chemical shifts of aro- matic residues in 13 C-labeled proteins via scalar couplings. J Am Chem Soc 1993;115:11054–11055. 56. Zwahlen C, Legault P, Vincent SJF, Greenblatt J, Konrat R, Kay LE. Methods for measurement of intermolecular NOEs by multi- nuclear NMR spectroscopy: application to a bateriophage k N- Peptide/boxB RNA complex. J Am Chem Soc 1997;119:6711– 6721. 57. Ottiger M, Bax A. Bicelle-based liquid crystals for NMR-measure- ment of dipolar couplings at acidic and basic pH values. J Biomol NMR 1999;13:187–191. 58. Tjandra N, Bax A. Direct measurement of distances and angles in biomolecules by NMR in a dilute liquid crystalline medium. Science 1997;278:1111–1114. Conformational Distribution for Apo-S100 Proteins PROTEINS 41
  • 15. 59. Hansen MR, Mueller L, Pardi A. Tunable alignment of macromole- cules by filamentous phage yields dipolar coupling interactions. Nat Struct Biol 1998;5:1065–1074. 60. Delaglio F, Grzesiek S, Vuister GW, Zhu G, Pfeifer J, Bax A. NMRPipe: a multidimensional spectral processing system based on UNIX pipes. J Biomol NMR 1995;6:277–293. 61. Johnson BA, Belvins RA. NMRView: a computer program for the visualization and analysis of NMR data. J Biomol NMR 1994;4: 603–614. 62. Brunger AT, Adams PD, Clore GM, DeLano WL, Gros P, Grosse- Kunstleve RW, Jiang JS, Kuszewski J, Nilges M, Pannu NS, Read RJ, Rice LM, Simonson T, Warren GL. Crystallography & NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr D Biol Crystallogr 1998;54(Part 5):905–921. 63. Gagne SM, Tsuda S, Li MX, Chandra M, Smillie LB, Sykes BD. Quantification of the calcium-induced secondary structural changes in the regulatory domain of troponin-C. Protein Sci 1994;3:1961– 1974. 64. Cornilescu G, Delagio F, Bax A. Protein backbone angle restraints from searching a database for chemical shift and sequence homol- ogy. J Biomol NMR 1999;13:289–302. 65. Nederveen AJ, Doreleijers JF, Vranken W, Miller Z, Spronk CA, Nabuurs SB, Guentert P, Livny M, Markley JL, Nilges M, Ulrich EL, Kaptein R, Bonvin JJ. RECOORD: a recalculated coordinate data- base of 5001 proteins from that PDB using restraints from that BioMagResBank. Proteins 2005;59:662–672. 66. Koradi R, Billeter M, Wuthrich K. MOLMOL: a program for dis- play and analysis of macromolecular structures. J Mol Graphics 1996;14:51–55. 67. Willard L, Ranjan A, Zhang H, Monzavi H, Boyko RF, Sykes BD, Wishart DS. VADAR: a web server for quantitative evaluation of protein structure quality. Nucleic Acids Res 2003;31:3316–3319. 68. Yap KL, Ames JB, Swindells MB, Ikura M. Vector geometry mapping. A method to characterize the conformation of helix-loop-helix cal- cium-binding proteins. Methods Mol Biol 2002;173:317–324. 69. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin TE. UCSF Chimera—a visualization system for ex- ploratory research and analysis. J Comput Chem 2004;25:1605–1612. 70. Lee YC, Volk DE, Thiviyanathan V, Kleerekoper Q, Gribenko AV, Zhang S, Gorenstein DG, Makhatadze GI, Luxon BA. NMR structure of the Apo-S100P protein. J Biomol NMR 2004;29:399– 402. 71. Fritz G, Mittl PRE, Vasak M, Grutters MG, Heizmann CW. The crystal structure of metal-free human EF-hand protein S100A3 at 1.7-A resolution. J Biol Chem 2002;277:33092–33098. 72. Tsuda S, Miura A, Gagne SM, Spyracopoulos L, Sykes BD. Low- temperature-induced structural changes in the Apo regulatory do- main of skeletal muscle troponin C. Biochemistry 1999;38:5693– 5700. 73. Barber KR, McClintock KA, Jamieson GA, Jr, Dimlich RV, Shaw GS. Specificity and Zn21 enhancement of the S100B binding epi- tope TRTK-12. J Biol Chem 1999;274:1502–1508. 74. Smith SP, Shaw GS. Assignment and secondary structure of cal- cium-bound human S100B. J Biomol NMR 1997;10:77–88. 75. Baxter NJ, Williamson MP. Temperature dependence of 1H chemi- cal shifts in proteins. J Biomol NMR 1997;9:359–369. 76. Smith SP, Barber KR, Shaw GS. Identification and structural influ- ence of a differentially modified N-terminal methionine in human S100b. Protein Sci 1997;6:1110–1113. 77. Inman KG, Baldisseri DM, Miller KE, Weber DJ. Backbone dynam- ics of the calcium-signaling protein apo-S100B as determined by 15N NMR relaxation. Biochemistry 2001;40:3439–3448. 78. Gagne SM, Tsuda S, Spyrocopoulos L, Kay LE, Sykes BD. Backbone and methyl dynamics of the regulatory domain of troponin C: ani- sotropic rotational diffusion and contribution of conformational entropy to calcium affinity. J Mol Biol 1998;278:667–686. 79. Spyracopoulos L, Lavigne P, Crump MP, Gagne SM, Kay CM, Sykes BD. Temperature dependence of dynamics and thermodynamics of the regulatory domain of human cardiac troponin C. Biochemistry 2001;40:12541–12551. 80. Babini E, Bertini I, Capozzi F, Luchinat C, Quattrone A, Turano M. Principal component analysis of the conformational freedom within the EF-hand superfamily. J Proteome Res 2005;4:1961–1971. 81. Notredame C, Higgins DG, Heringa J. T-Coffee: a novel method for fast and accurate multiple sequence alignment. J Mol Biol 2000;302: 205–217. 82. Zhang M, Tanaka T, Ikura M. Calcium-induced conformational transition revealed by the solution structure of apo calmodulin. Nature Struct Biol 1995;2:758–767. 83. Gagne SM, Li MX, McKay RT, Sykes BD. The NMR angle on tro- ponin C. Biochem Cell Biol 1998;76:302–312. 84. Gagne SM, Tsuda S, Li MX, Smillie LB, Sykes BD. Structures of the troponin C regulatory domains in the apo and calcium-saturated states. Nat Struct Biol 1995;2:784–789. 85. Bhattacharya S, Bunick CG, Chazin WJ. Target selectivity in EF- hand calcium binding proteins. Biochim Biophys Acta 2004;1742: 69–79. 86. Maler L, Sastry M, Chazin WJ. A structural basis for S100 protein specificity derived from comparative analysis of apo and Ca21 -cal- cyclin. J Mol Biol 2002;317:279–290. S. Malik et al. 42 PROTEINS