You are on page 1of 11

View metadata, citation and similar papers at core.ac.

uk brought to you by CORE


provided by Elsevier - Publisher Connector
Cell, Vol. 122, 183–193, July 29, 2005, Copyright ©2005 by Elsevier Inc. DOI 10.1016/j.cell.2005.05.033

Structural Basis for the EBA-175


Erythrocyte Invasion Pathway
of the Malaria Parasite Plasmodium falciparum
Niraj H. Tolia,1,2 Eric J. Enemark,2 B. Kim Lee Sim,3,* velopment of interventions to control malaria. Several
and Leemor Joshua-Tor1,2,* specific ligand-receptor interactions have been iden-
1
Watson School of Biological Sciences tified for invasion and genetic disruption of any one in-
2
Keck Structural Biology Laboratory teraction results in a shift to using other pathways (Dur-
Cold Spring Harbor Laboratory aisingh et al., 2003a, 2003b; Reed et al., 2000; Triglia et
1 Bungtown Road al., 2005).
Cold Spring Harbor, New York 11724 EBA-175 is a P. falciparum protein that binds its re-
3
Protein Potential ceptor glycophorin A (GpA) on human erythrocytes dur-
9640 Medical Center Drive ing invasion (Adams et al., 1992; Camus and Hadley,
Rockville, Maryland 20878 1985; Klotz et al., 1992; Orlandi et al., 1992; Sim et al.,
1990). GpA is the major glycoprotein found on human
erythrocytes and is heavily sialylated (Marchesi et al.,
Summary 1972). The EBA-175/GpA pathway is the dominant chy-
motrypsin-resistant invasion pathway (Duraisingh et al.,
Erythrocyte binding antigen 175 (EBA-175) is a P. fal- 2003a). Antibodies against EBA-175 inhibit binding to
ciparum protein that binds the major glycoprotein GpA and block invasion of merozoites in vitro (Sim et
found on human erythrocytes, glycophorin A, during al., 1990, 2001). The receptor-ligand interaction in-
invasion. Here we present the crystal structure of the volves sialic acids on the mucin domain of GpA, prefer-
erythrocyte binding domain of EBA-175, RII, which entially Neu5Ac(α2-)Gal(Neu5Ac(α-2,6)Gal) on O-linked tet-
has been established as a vaccine candidate. Binding rasaccharides. GpA, the major sialoglycoprotein found on
sites for the heavily sialylated receptor glycophorin A erythrocytes, is a 131 amino acid transmembrane dimer.
are proposed based on a complex of RII with a glycan Each monomer spans the membrane once exposing its
that contains the essential components required for N terminus extracellularly (MacKenzie et al., 1997). The
binding. The dimeric organization of RII displays two erythrocyte binding domain of EBA-175 is region II (RII)
prominent channels that contain four of the six ob- (Figure 1). RII is a 616 amino acid fragment of EBA-175
served glycan binding sites. Each monomer consists that consists of two cysteine-rich regions—F1 and F2.
of two Duffy binding-like (DBL) domains (F1 and F2). F1 and F2 are homologous to the Duffy binding protein
F2 more prominently lines the channels and makes of P. vivax and are therefore called Duffy binding-like
the majority of the glycan contacts, underscoring its (DBL) domains. The presence of one or two DBL do-
role in cytoadherence and in antigenic variation in mains, a C-terminal cysteine-rich region (region VI), and
malaria. Our studies provide insight into the mecha- a type I transmembrane-like domain define EBA-175 as
nism of erythrocyte invasion by the malaria parasite a member of the erythrocyte binding-like (EBL) super-
and aid in rational drug design and vaccines. family (Adams et al., 1992; Figure 1). The EBL super-
family includes several other proteins believed to be
Introduction involved in erythrocyte invasion, such as the following:
Duffy binding protein, BAEBL/EBA-140/EBP2 (Mayer et
Malaria causes an estimated 300–500 million clinical al., 2001; Narum et al., 2002; Thompson et al., 2001),
cases and 1–3 million deaths annually (Breman et al., EBA-181/JESEBL (Gilberger et al., 2003b), and MAEBL
2001). More than 80% of deaths from malaria occur in (Blair et al., 2002), and other candidate genes where
children in sub-Saharan Africa (Snow et al., 1999). The the presence of protein was not identified (Triglia et al.,
disease is caused by parasites of the genus Plasmo- 2001). Also included are the var genes, which are a
dium, of which Plasmodium falciparum is the most large family of genes regulated by chromatin modifica-
prevalent and is responsible for almost all the associ- tion (Duraisingh et al., 2005; Freitas-Junior et al., 2005)
ated mortality. Sporozoite-stage malarial parasites in- that encode antigenically variant proteins including
oculated by an infected female Anopheles mosquito PfEMP-1 (Baruch et al., 1995; Smith et al., 1995; Su
rapidly invade hepatocytes. After five to six days, the et al., 1995). PfEMP-1 contains several copies of DBL
parasites become mature liver-stage schizonts with domains and is responsible for cytoadherence of para-
thousands of merozoites. When the hepatocyte rup- sitized erythrocytes to endothelium causing the most
tures, the merozoites are released into the blood fatal form of malaria. The signature DBL domains of the
stream. Each of these merozoites can invade an eryth- EBL family have been found to be important for recep-
rocyte, initiating the erythrocytic stage of the parasite’s tor binding (Mayor et al., 2004) and it appears that these
life cycle (for review see Miller et al. [2002]). domains are unique to Plasmodium and thus constitute
All of the pathological and clinical manifestations of a good drug target (Aravind et al., 2003).
the disease are caused by the asexual erythrocytic RII is highly conserved among more than 30 labora-
stage. Thus, the invasion of erythrocytes by merozoites tory and field isolates (Liang and Sim, 1997) and has
marks a critical step that is a logical target for the de- been established as a vaccine candidate (Jones et al.,
2001; Sim et al., 2001). Toward the understanding of
*Correspondence: leemor@cshl.edu (L.J.); ksim@protpot.com receptor-ligand interactions critical for erythrocyte in-
(B.K.L.S.) vasion by P. falciparum, we have determined the crystal
Cell
184

Figure 1. DBL-Containing Proteins from Plasmodium Species


(A) The ebl family shares a conserved gene structure that encodes either one or two DBL domains in region II. The var superfamily contains
genes that encode multiple repeats of DBL domains. MAEBL is a unique member of the EBL family. It has the conserved domain structure,
but its region II does not contain DBL domains. Instead, the MAEBL ligand domains are related to another merozoite protein, apical membrane
antigen 1 (AMA-1). PvDBP—Plasmodium vivax Duffy binding protein; PkDBP—Plasmodium knowlesi DBP; PfEMP-1—Plasmodium falciparum
erythrocyte membrane protein 1.
(B) Alignment and structural depiction of DBL-containing proteins. Light blue indicates conserved residues, and deep blue indicates invariant
residues. Cysteines are shown in red. Secondary structure is shown above the sequence alignment with F1 in green, linker regions in gold,
and F2 in purple. Dashed lines indicate regions not observed in the structure of RII. Blue and red dots indicate residues involved in dimeriza-
tion and glycan binding, respectively.
Crystal Structure of a Malaria Vaccine Candidate
185

structure of EBA-175 RII of P. falciparum strain 3D7 to


2.3 Å resolution.
The structure shows a dimeric form of RII, with exten-
sive interactions between the elongated monomers.
Two channels are prominent in the center of the dimer.
The F1 and F2 domains are mostly helical and are sim-
ilar in structure. To identify possible receptor binding
sites, we cocrystallized RII with α-2,3-sialyllactose. The
glycan binding sites suggest a possible interaction sur-
face for the receptor. Our structural studies facilitate
the elucidation of the mechanism of erythrocyte inva-
sion and allow the rational design of receptor blockade
therapy and vaccines.

Results and Discussion

Overall Structure of RII


The unbound RII domain of EBA-175 of P. falciparum
crystallized in two distinct crystal forms. The structure
was determined by multiwavelength anomalous disper-
sion (MAD) (see Experimental Procedures). The overall
organization of RII in both crystal forms is conserved,
with an rmsd for all Cα atoms of 0.9 Å.
The crystal structure of the RII monomer reveals an
elongated molecule, composed primarily of α helices,
with two antiparallel β hairpins and several bound sul-
fate molecules (Figure 2A). The two DBLs, F1 (residues
8–282) and F2 (residues 297–603), have a very similar
structure with an rmsd of 1.96 Å (calculated by LSQMAN
[Kleywegt, 1999] for 194 Cα atoms; Figure 2B). The two
domains are connected by a three helix linker (283–
317). These domains possess a novel fold, based on
an analysis with DALI (Holm and Sander, 1993). Each
domain is composed of two subdomains. The topology
of RII, shown in Figure 2C, demonstrates the similarity
of the two repeats and depicts the two subdomains of
each repeat. All but one of the cysteines, C273, are in-
volved in disulfide bridges. C598 is not well ordered in
the structure of unbound RII but is in a clear disulfide
bridge with C513 as seen in the complex described be-
low. All disulfide bridges are contained within a given
subdomain (Figure 2C). The bridges superimpose nearly Figure 2. The RII Monomer
perfectly between the two domains, and stabilize the (A) Overall structure of the RII monomer. F1—green; F2—purple;
RII cysteine-rich repeat fold (F1/F2). linker—gold; bound sulfates—CPK spheres.
Alignment of EBA-175 RII with other RIIs from the (B) Overlay of RII subdomains F1 (green) and F2 (purple). Red
bonds and yellow bonds indicate disulfide bridges in F1 and F2, re-
EBL family of proteins (Figure 1B) reveals 34 invariant spectively.
residues. With the exception of Q43, W330, W450, W458 (C) Topology diagram of RII denoting similarity between F1 (green)
and W528, all invariant residues are involved in fold sta- and F2 (purple). Disulfide bridges (red) are all within subdomains of
bilization, further contributing to the stability of the DBL F1 and F2. The dotted line indicates portions of the structure or-
domain fold. As with the disulfide bridges, a large num- dered in the complex but not in unbound RII.
ber of the invariant residues interact almost exclusively
with other residues in the same subdomain. Exceptions
are a salt bridge between R349 and E488, hydrogen with the F2 domain of the second monomer and vice-
bonds (H bonds) between D414 and Q481, and hy- versa. This dimeric interaction is conserved in all crys-
drophobic interactions between W330 and W489 and tal forms of RII. The buried surface area is 1480 Å2 per
between R349 and W485, all of which are involved in dimer, consistent with the expected size of a total bur-
interactions between the two subdomains of F2. ied area of 1600 ± 400 Å2 for noncovalently interacting
proteins (Lo Conte et al., 1999). Dynamic light scatter-
The RII Dimer—a Handshake Interaction ing (DLS) experiments show that RII exists as a dimer
RII crystallizes as a dimer where two symmetrically re- with a Stoke’s radius of 4.29 ± 0.87 nm and an apparent
lated RII molecules interact extensively with each other molecular weight of 160 kDa at high concentrations
in an antiparallel fashion resembling a handshake (Fig- (w2 mg/ml; data not shown). However, analytical ultra-
ure 3A). The F1 domain of one monomer interacts mostly centrifugation experiments at lower concentrations (0.4
Cell
186

Figure 3. The RII Dimer


(A) Stereo ribbon representation of the RII di-
mer. Monomers are shaded in different inten-
sities (light/dark) for clarity. F1 domains are
green or light green; F2 domains are purple
or light purple, and the linker regions are
gold or light gold. Sulfates are depicted in
ball and stick (sulfurs in yellow, oxygens in
red).
(B) Two views of the electrostatic surface po-
tential of the RII dimer calculated by GRASP
at an ionic strength of 0.1 M. Surface poten-
tial is colored is from −10 kT (red) to +10 kT
(blue). Sulfates can be clearly seen on the
“bottom” surface on the left and in the
channels.
(C) A close up view of the β finger of F1
(green) inserted into a cavity of F2 (purple).
(D) A close up view of the β finger of F2 (pur-
ple) and the cavity of F1 (green).
(E) Interactions between the two-stranded
antiparallel β sheet of the two F2s. Coloring
scheme as in (A). Dashed lines indicate hy-
drogen bonds and/or electrostatic interac-
tions.

mg/ml) indicate RII is predominantly monomeric with a ules we termed “β fingers” that mediate intermolecular
small percentage of dimers (data not shown). In addi- interactions. Residues of the F1 β finger insert into a
tion, crosslinking studies with bis-sulfosuccinimidyl cavity in F2 (Figure 3C), and residues of the F2 β finger
show that RII exists in both monomeric and dimeric insert into a cavity in F1 (Figure 3D). Some key interac-
forms (data not shown). Thus, we observe a mixture tions are a salt bridge between D30 (F1 β finger) and
of both species, monomers and dimers, in solution. A R446 (F2 cavity) and van der Waals interactions be-
surface representation of the RII dimer (Figure 3B) tween T114 (F1 cavity) and L338 and T340 (F2 β finger).
shows a highly basic protein surface (the calculated pI The dimer is further stabilized by H bonds between the
is 8.8) and reveals the presence of two channels that N118 side chain (F1 cavity) and the backbone carbonyl
span the dimer. The channel surfaces consist of resi- of S339 of the F2 β finger and between the carbonyl of
dues from both monomers. Most of these residues, T340 (F2 β finger) and the amide group of Q121 (F1
about two-thirds, come from the two F2 domains of cavity). The third intermolecular interaction is the for-
the dimer. mation of a two strand antiparallel β sheet between
The two β hairpins of F1 and F2 form structural mod- identical F2 residues (N433-H436—β5) of each mono-
Crystal Structure of a Malaria Vaccine Candidate
187

mer (Figure 3E). This β sheet is found in the center of Functional Analysis of RII Mutants
the dimer, and separates the two channels. To test the relevance of the dimeric state and glycan
Several well-ordered sulfate molecules are observed binding sites to RII erythrocyte binding, we performed
at the same positions in both crystal forms. Interest- a functional analysis on selected residues involved in
ingly, these are bound exclusively in the channels and these characteristics of RII. RII single point mutants
on one face of the dimer (Figure 3B), which for our dis- were individually expressed on the surface of COS cells
cussion we call the “bottom” surface. and tested for their ability to bind normal human eryth-
rocytes (Table 1; Figure S3; Sim et al., 1994).
Residues involved in the dimeric interaction include
Receptor Binding Sites D30 and R446, which form a direct salt bridge as well
In order to locate receptor binding surfaces and study as a water-mediated interaction between the mono-
the binding mode of RII to the GpA receptor, we cocrys- mers (Figure 3C); T114, which forms van der Waals in-
tallized RII with the glycan α-2,3-sialyllactose. This gly- teractions with side chains of residues L338 and T340
can contains the essential component, Neu5Ac(α2,3)- (Figure 3D); and backbone atoms in N433-H436 of both
Gal, of the receptor glycan that is required for binding monomers that form an antiparallel β sheet (Figure 3E).
(Neu5Ac denotes sialic acid). The complex crystallized To disrupt the D30/R446 salt bridge, R446 was mutated
in a third, related crystal form, with one dimer in the to a glutamate (R446E) and an aspartate (R446D). Both
asymmetric unit (see Experimental Procedures). The mutants demonstrated reduced binding to erythrocytes
protein structure is largely unchanged (rmsd 0.39 Å), (Table 1). Steric disruption of interactions with T114 by
though there is a slight reduction in the hinge angle mutation to phenylalanine (T114F) resulted in a reduc-
between F1 and F2 in the complex. In addition, resi- tion in binding efficiency (Table 1). Since the interac-
dues 597–601 that were disordered in the unbound tions that form the antiparallel β sheet in the center of
structure are ordered in the complex. This segment in- the dimer are all backbone interactions, V435 was mu-
cludes C598, which is now clearly seen in a disulfide tated to aspartate (V435D) to introduce charge repul-
bond with C513. All but two of the sulfates are ob- sion. Again, a reduction in binding is observed (Table
served here as well. 1). Therefore, we conclude that RII dimerization is im-
Six glycan binding sites are observed in the RII dimer, portant for erythrocyte binding.
all at the dimer interface (Figure 4 and see Figure S1 in Several residues involved in glycan binding were also
the Supplemental Data available with this article on- selected for mutagenesis. N417, R422, and K439 are
line). Four of these, 1–4, are inside the channels. The involved in interactions with glycans 1 and 2. Mutation
other two, 5 and 6, are exposed to the top surface and of these residues result in reduced binding (Table 1).
are separated from the channel by the β fingers and by Note that, unlike the others, R422 is a severe mutation
glycans 1 and 2. There is also a sulfate between glycan to glutamate, which may explain its stronger effect.
positions 1 and 5 and between 2 and 6. Glycan posi- N33, N551, Y552, and K553, which are all involved in
tions 1 and 2, 3 and 4, and 5 and 6 are related by an interactions with glycans 3 and 4, were mutated to ala-
approximate 2-fold axis. Though it is possible to model nine. Each reduced binding to different extents. Finally,
sialic acid moieties for all glycans, and even an addi- K28, R31, and K341 interact with glycans 5 and 6, and
tional galactose moiety for glycan positions 5 and 6 mutation of these residues to alanine also resulted in
(Figure S1), it was not possible to model these accu- varying degrees of reduced binding. The relative levels
of reduction in binding roughly correlate with the num-
rately and unambiguously due to somewhat limited res-
ber of glycan contacts made by each residue (Tables 1
olution (2.25 Å) combined with partial occupancy (0.4–
and S1; Figure S1D).
0.6) for the glycans, resulting in a mixture of bound and
This analysis suggests that all six glycan binding
unbound conformations for some of the side chains in
sites are important for RII binding to GpA, as mutations
the vicinity of these glycans.
that are predicted to disrupt binding to the glycans ob-
All six glycans contact both monomers, implicating
served in the structure of the complex prevent binding
dimerization as important for receptor binding (Figures
of RII to red blood cells. The reduction or lack of bind-
4B and S1D; Table S1). Glycans 1 and 2 contact resi-
ing by these mutants was not due to inappropriate ex-
dues N417, R422, N429, K439, and D442 of one mono-
pression, as expression levels were comparable to wild-
mer and K28 of the other monomer. It should be noted type RII as shown by immunofluorescence (Figure S3). As
that Y55 from a neighboring complex also contacts gly- a negative control, two residues on the outer surface of
cans 1 and 2; however, since this is a lattice interaction, the RII dimer that are not predicted to be involved in
its significance is not clear. Glycans 3 and 4 contact dimerization or glycan binding were changed to ala-
residues N550, N551, Y552, K553, and M554 from one nine, and these show full binding activity (Table 1).
monomer and N33 from the other. Glycans 5 and 6 con-
tact residues T340, K341, D342, V343, Y415, Q542, and A Model for Erythrocyte Binding
Y546 of one monomer and residues K28, N29, R31, and GpA exists as an erythrocyte transmembrane dimer
S32 of the second monomer. We modeled the full with two heavily glycosylated extracellular domains per
O-glycan of GpA, Neu5Ac(α2-3)Gal (NeuAc(α-2,6)Gal), dimer. The crystal structure of RII, the subject of this
at each of the glycan binding sites observed in the study, reveals a dimer with two channels. In addition,
complex to test whether it would fit. Indeed these we identified six glycan binding sites per dimer of RII,
studies indicate that the full O-glycan could be accom- four of which (1–4) are accessible from the channels
modated in all of these sites with good geometry (Fig- and the other two (5 and 6) are accessible through a
ure S2). cavity on the “top” surface of RII.
Cell
188

Figure 4. Crystal Structure of RII with Sialyllactose


(A) Ribbon diagram of RII with the position of the glycans indicated by the Fo − Fc electron density map contoured at 2.5 σ in red.
(B) Close-up view of three of the glycan binding sites, glycans 1, 3, and 5 are shown from left to right. All glycans are contacted by residues
from both monomers. Color scheme as in Figure 3.

The extracellular domain of GpA inhibits RII binding suggests a mode for GpA binding (Figure 5). Since the
to erythrocytes, implicating this region of GpA as the glycans are bound at the dimer interface and make
receptor for RII (Sim et al., 1994). The presence of two contacts to both monomers, we reason that the RII di-
channels in the center of the RII dimer along with the mer might assemble around the dimeric GpA extracel-
location of the glycans in the sialyllactose complex lular domain and that dimerization of RII may occur
upon ligand-receptor binding at the erythrocyte sur-
face during the invasion process. Moreover, binding-
Table 1. Binding Efficiency for RII and Mutants induced dimerization is consistent with our observation
that glycan 5/6 is bound in a deep elongated pocket
Binding No. Glycan
Mutant Class Mutation Ability Contacts
created by both monomers. Although this pocket is ac-
cessible through a cavity on the top surface of the di-
Wildtype RII ++++ N/A
mer, the glycan is too bulky to enter without some
Dimer R446E — N/A
Dimer R446D — N/A
movement of the protein. On the other hand, GpA bind-
Dimer T114F + N/A ing is straightforward if monomers would assemble
Dimer F2 V435D ++ N/A around it. Thus, any preexisting dimers would dissoci-
Glycan 1/2 N417A ++ 2 ate to allow for receptor binding. This scenario is con-
Glycan 1/2 R422E — 3 sistent with both analytical ultracentrifugation and
Glycan 1/2 K439A + 2 crosslinking studies, mentioned above, that indicate
Glycan 3/4 N33A + 1
the predominance of monomers in solution. In addition,
Glycan 3/4 N551A +++ 1
Glycan 3/4 Y552A + 2
we should note that in a separate study, recombinant
Glycan 3/4 K553A ++ 2 non-N-glycosylated RII in phosphate buffered saline at
Glycan 5/6 K28A +++ 2 pH 7.4 manufactured for clinical use is a stable mono-
Glycan 5/6 R31A ++ 1 mer (B.K.L.S., unpublished data). The remaining do-
Glycan 5/6 K341A + 3 mains of EBA-175 may also influence the assembly of
Surface control E153A ++++ N/A RII around GpA, impacting a conversion from monomer
Surface control K548A ++++ N/A
to dimer. Invasion involves dramatic changes in the
Crystal Structure of a Malaria Vaccine Candidate
189

Figure 5. Model for P. falciparum EBA-175-


RII Binding to the Red Blood Cell Receptor
Glycophorin A (GpA)
The P. falciparum membrane is shown on the
top and the erythrocyte membrane on the
bottom. The receptor binding domain of
EBA-175, RII, is shown as a surface repre-
sentation with F1 in green and F2 in purple
and the linker in gray. Blue lines represent
portions of EBA-175 backbone not included
in the crystal structure. GpA is shown in red
with the membrane-spanning region in detail
using the NMR structure and the extracellu-
lar domain drawn as a schematic flexible
line. The modeled O-glycans of GpA (see
Figure S2) are shown as space-filling models
in gold. In the left panel, the RII dimer as-
sembles around the GpA dimer, with GpA
binding within the channels (see text). An al-
ternative model is shown on the right, where
the GpA monomers dock on the outer sur-
face of the protein, feeding glycans into the
channels.

parasite that would require signaling within the merozo- ization interface of the cavity of F2 and includes R446,
ite. It has been shown that the cytoplasmic C-terminal which we show here to be required for binding to eryth-
domain of EBA-175 is essential for invasion but not for rocytes. This peptide also includes residues that line
trafficking of the protein (Gilberger et al., 2003a). This the channels of the dimer and are involved in glycan
suggests that signaling occurs through the cytoplasmic interactions. Thus, peptide 435–455 may function either
domain during the invasion process and that this sig- by binding to GpA directly, thereby capping GpA on
naling might be triggered following dimerization upon erythrocytes and blocking merozoite invasion, or by
ligand binding. binding to EBA-175 RII preventing dimerization. Muta-
The channels are 15 Å in diameter at their narrowest; genesis of this peptide also led to the generation of
therefore, this region of the channel would be able to more immunogenic peptides that protect from chal-
accommodate an unglycosylated segment of GpA, lenge with P. falciparum (Guzman et al., 2002).
such as residues 5–9, 16–21, 27–36, or 38–43. The gly-
cosylated regions of GpA would be bound to the outer Implications for the DBL Superfamily
segments of the channels and the top face of RII, while Several erythrocyte invasion pathways have been iden-
an unglycosylated region of GpA would be bound at tified in P. falciparum. Some pathways involve DBL-
the center, the narrowest part, of the channel (Figure 5). containing proteins, in particular those of the EBL fam-
Alternatively, the GpA monomers may dock on the ily, with the EBA-175/GpA pathway being the dominant
outer surface of RII, feeding glycans into the channels chymotrypsin-resistant invasion pathway (Duraisingh et
(Figure 5). Since the structure indicates that the glycan al., 2003a). Other pathways do not involve DBL-con-
binding sites are formed by both monomers, dimeriza- taining proteins, such as the reticulocyte binding-like
tion of RII would greatly enhance binding to GpA. In- family (Duraisingh et al., 2003b; Kaneko et al., 2002;
deed, mutations of residues involved in dimerization or Rayner et al., 2000; Rayner et al., 2001; Taylor et al.,
in putative glycan binding impair the ability of RII to 2002; Triglia et al., 2005). Apart from the cysteine-rich
bind erythrocytes (Table 1). binding domain proteins of P. falciparum, P. vivax, and
It was shown previously that F2 alone can bind eryth- P. knowlesi, involved in erythrocyte invasion, the DBL
rocytes, but that F1 is insufficient (Sim et al., 1994). F2 superfamily also includes the highly polymorphic
contributes about two-thirds of the residues that line multicopied var gene products adapted for functions
the channels (see Figures 3A and S2). In addition, there including adhesion and receptor recognition. The struc-
are dimeric interactions exclusively between F2 mono- ture of EBA-175 RII now allows modeling of these DBL-
mers, but F1 depends on F2 for dimerization (Figures containing proteins and predictions about the alternate
3E and S2). Thus, we speculate that F2 alone can form DBL-dependent pathways of invasion. Furthermore,
a weak dimer or can be induced to dimerize by the re- models of the DBL domains of PfEMP-1 could reveal
ceptor. Moreover, F2 also contributes 75% of the con- conserved regions in the variant PfEMP-1 DBL domains
tacts to the glycans observed in the complex with sial- and/or lead to an understanding of PfEMP-1 receptor/
yllactose described above. Thus it appears to provide ligand interactions. Figure 1B shows a sequence align-
most of the interaction sites for the receptor. ment of the double DBL domain containing RIIs of
Interestingly, two peptides (355–375 and 435–455) EBA-175, BAEBL, EBL-1, and JESEBL. Interestingly,
generated from RII that were shown to bind erythro- the sequence conservation is higher in F2 than in F1,
cytes and inhibit in vitro merozoite invasion (Rodriguez suggesting a greater selective pressure on F2. This is
et al., 2000) mapped to residues lining the channels in accordance with the observation that binding is de-
(Figure S4). Peptide 435–455 includes part of the dimer- pendent on F2 to a greater extent than F1. Most invari-
Cell
190

ant residues including the cysteines are involved in methylotropic yeast. Fermentation was performed with defined
interactions that stabilize the DBL domain fold. In addi- salt media.
Purification was initiated by expanded bed adsorption chroma-
tion, a large number of residues that are involved in
tography, i.e., sulphopropyl cation exchange chromatography (Am-
Neu5Ac(α2-3)Gal binding are conserved among five ersham Pharmacia Biotech), followed by additional ionic exchange
RIIs of the EBL family. chromatographic steps (SP chromatography Streamline SP XL
It is interesting that F1 is similar to the single DBL resin and Q FF [Pharmacia]). The final steps of the purification in-
domain of P. vivax and P. knowlesi Duffy binding pro- cluded a Streamline SP/XL cation exchange polishing step and an
teins that have been shown to bind erythrocytes during ultrafiltration diafiltration concentration step.
invasion (Adams et al., 1992; Chitnis and Miller, 1994).
There are also similarities in amino acid signatures be- Protein Crystallization and Derivatization
tween F1 and the DBL and CIDR domain of the var Form 1 crystals were grown by vapor diffusion by mixing 2 ␮l of
the protein solution (15 mg/ml in 10mM Tris-HCl [pH 7.4], 100mM
gene family that encode proteins responsible for cy-
NaCl) with 2 ␮l of a reservoir solution containing 0.265 M ammo-
toadherence and antigenic variation. Thus, the three- nium sulfate, 0.1 M sodium cacodylate (pH 6.5), 29% polyethylene
dimensional structure of the RII DBL domain should be glycol 8000. Form 2 crystals were grown similarly but with a reser-
conserved throughout the DBL protein family. The ob- voir containing 2.6–2.8 M ammonium sulfate, 0.1 M sodium caco-
served amino acid differences may be driven by pres- dylate (pH 6.5), 0.05%–2% polyethylene glycol 750 monomethyl
sure for immune evasion. Indeed, in a previous study ether, and a protein concentration of 12 mg/ml. Derivatized crystals
were obtained by soaking form 1 or form 2 crystals in saturated
(Liang and Sim, 1997), a strong conservation of both
lithium sulfate containing either 1 mM sodium dicyanoaureate or
the structure and function of EBA-175 RII was found 1 mM potassium tetrachloroplatinate for 2 hr.
across laboratory strains and field isolates with only 16 Complex crystals were grown by mixing 2 ␮l of the protein solu-
point mutations leading to radical amino acid changes tion (20–30 mg/ml in 10 mM Tris-HCl [pH 7.4], 100 mM NaCl) with
that were few and scattered. Localization of the amino 0.4–1.0 ␮l of a 10 mM α-2,3-sialyllactose (Sigma) solution and 2 ␮l
acids that were involved in the point mutations revealed of a reservoir solution containing 2.7 M ammonium sulfate, 0.1 M
HEPES (pH 7.5), and 0.1% polyethylene glycol 750 monomethyl
that 75% of them were on the surface of RII arguing
ether.
that these mutations might have arisen from immune
pressure.
Data Collection
Regions of the EBL family that are surface exposed
Native and complex crystals were cryoprotected by transfer to a
could be used for targeted vaccine design. Loop re- solution of saturated lithium sulfate and flash frozen in liquid nitro-
gions are often good targets for use in immunization gen. Data were collected at beamlines X26C and X25 at the Na-
due to their inherent flexibility and conformability. Fur- tional Synchrotron Light Source (NSLS) at Brookhaven National
thermore, the use of residues involved in dimerization Laboratory (BNL). For derivatives, crystals were flash frozen di-
or in glycan binding may prove fruitful for targeted vac- rectly from the soak solution prior to data collection. Multiwave-
length anomalous dispersion data were collected for the gold and
cine design as described for peptide 435–455. Design
for the platinum derivatives. Data collection statistics are shown in
of small molecule inhibitors using the glycan binding Tables S2 (form 1) and S3 (form 2).
sites could prove useful as well. Though protein-protein Data reduction statistics for the complex indicated a loss of a
interactions including dimerization have not been tradi- crystallographic 2-fold had occurred upon ligand binding. This re-
tional targets for drug intervention, using a small mole- sulted in a conversion from a tetragonal P4122 to an orthorhombic
cule targeted to a well-formed cavity as seen in a crys- C2221 space group (see Table S4). The abolished 2-fold relates the
two monomers of the dimer, which is a noncrystallographic 2-fold
tal structure has proven successful, with disruption of
in the complex.
the p53-MDM2 interaction as a recent example (Vassi-
lev et al., 2004). The F2 cavity, required for dimerization,
which in turn is necessary for EBA-175 function, could Structure Solution and Analysis
Four gold sites were identified by SnB (Weeks and Miller, 1999) and
serve as such a target.
confirmed by SHELXS (Sheldrick et al., 1993) with data collected
Here, we have solved the crystal structure of a at the absorption peak for a form 1 gold derivative. Following phase
double-DBL domain from the EBL family of proteins un- refinement with SHARP (de la Fortelle and Bricogne, 1997), plati-
bound and in complex with α-2,3-sialyllactose. These num sites in form 1 were determined by crossphasing with CNS
structures provide a glimpse into the mechanism of (Brünger et al., 1998). Phase refinement with the form 1 native and
erythrocyte invasion by the Plasmodium parasite. They derivative data (4 Au wavelengths, 3 Pt wavelengths with SHARP)
allowed the building of a polyalanine model comprising w60% of
also provide insights into a large superfamily of pro-
the backbone in O (Jones and Kjeldgaard, 1997).
teins critical for the pathogenesis and survival of the This model was separated into two halves and used to solve the
parasite. Lastly, this study will help the rational design structure of form 2 by molecular replacement (AMoRe [Navaza and
of novel therapeutics that could aid in the desperately Saludjian, 1997]). Four low-occupancy gold sites were then iden-
needed treatment of malaria. tified by difference Fourier for a form 2 gold derivative. Simulta-
neous cross crystal averaging, density modification, and phase ex-
tension (DMMULTI–CCP4 [CCP4, 1994]) using both form 1 and 2
Experimental Procedures native data and form 1 and 2 initial phases improved the density to
allow w90% of the backbone and 60% of sequence to be assigned
Protein Expression and Purification unambiguously. Iterative SIGMAA (CCP4) weighting, RESOLVE (Ter-
The native RII domain of EBA-175 is unglycosylated in P. falci- willliger, 2000) prime-and-switch density modification, and model
parum due to the lack of posttranslational modification capabilities building (O) allowed unambiguous assignment of 95% of the se-
for glycosylation (Gowda and Davidson, 1999); however, expres- quence. Phasing statistics are shown in Tables S2 and S3.
sion in Pichia pastoris resulted in a glycosylated protein that was Refinement of the model and rebuilding was performed with
refractory to crystallization. A synthetic gene fragment encoding RII CNS, Reduce and Probe (Word et al., 1999), MolProbity (Lovell et
with five putative glycosylation sites mutated (N3Q, S50A, S195A, al., 2003), and O. Final refinement statistics are shown in Table S4.
T206A, N400Q) was cloned and expressed in Pichia pastoris, a The final model contains residues 8–163, 166–508, and 513–596 of
Crystal Structure of a Malaria Vaccine Candidate
191

RII, 326 water molecules, 9 sulfates, and one chloride ion with an Received: March 2, 2005
Rwork/Rfree of 0.232/0.280. Revised: April 22, 2005
The sialyllactose complex was solved by molecular replacement Accepted: May 24, 2005
(AMoRe) using the refined form 2 structure without the solvent Published: July 28, 2005
molecules as a search model. Two monomers were found in the
asymmetric unit. Initially, TLS and tightly restrained NCS refinement References
with REFMAC5, was followed by model rebuilding and the addition
of solvent molecules with O. The locations of the six glycans were Adams, J.H., Sim, B.K., Dolan, S.A., Fang, X., Kaslow, D.C., and
identified by visual inspection of Fo − Fc electron density maps, Miller, L.H. (1992). A family of erythrocyte binding proteins of ma-
calculated with CNS and shown in Figure S1A. Final refinement laria parasites. Proc. Natl. Acad. Sci. USA 89, 7085–7089.
statistics are shown in Table S4. This protocol was chosen based Aravind, L., Iyer, L.M., Wellems, T.E., and Miller, L.H. (2003). Plas-
on careful monitoring of the Rfree. The final model contains two modium biology: genomic gleanings. Cell 115, 771–785.
monomers of RII each composed of residues 8–163, 166–508, and
Bacon, D.J., and Anderson, W.F. (1988). A fast algorithm for render-
513–601, 599 water molecules, 16 sulfates, and two chlorides with
ing space-filling molecule pictures. J. Mol. Graph. 6, 219–220.
an Rwork/Rfree of 0.216/0.232. A possible model for the glycans was
also built and the resulting 2Fo − Fc electron density maps for the Baruch, D.I., Pasloske, B.L., Singh, H.B., Bi, X., Ma, X.C., Feldman,
glycans are shown in Figure S1B. In addition, a simulated-annealed M., Taraschi, T.F., and Howard, R.J. (1995). Cloning the P. falci-
Fo − Fc omit map (Figure S1C) was calculated by omitting the gly- parum gene encoding PfEMP1, a malarial variant antigen and ad-
cans as well as a 2 Å sphere around them. Though the general herence receptor on the surface of parasitized human erythrocytes.
placement of these glycans was obvious, a precise atomic model Cell 82, 77–87.
was not possible due to the relatively limited resolution. In addition, Blair, P.L., Kappe, S.H., Maciel, J.E., Balu, B., and Adams, J.H.
due to only partial occupancy of the glycans (0.4–0.6), the protein (2002). Plasmodium falciparum MAEBL is a unique member of the
had to be modeled using alternate conformations for side chains ebl family. Mol. Biochem. Parasitol. 122, 35–44.
in bound and unbound states. Less-than-ideal density is not un- Breman, J.G., Egan, A., and Keusch, G.T. (2001). The intolerable
usual for carbohydrates at this resolution. burden of malaria: a new look at the numbers. Am. J. Trop. Med.
Hyg. 64, iv–vii.
Mutagenesis and Red Blood Cell Binding (Rosette) Assays Brünger, A.T., Adams, P.D., Clore, G.M., DeLano, W.L., Gros, P.,
Single amino acid mutations were introduced in RII, previously Gross-Kunstleve, R.W., Jiang, J.S., Kuszewski, J., Nilges, M.,
cloned into plasmid pRE4, by the Quickchange method (Strata- Pannu, N.S., et al. (1998). Crystallography & NMR system: a new
gene). Fresh monolayers of COS-7 cells cultured on coverslips in software suite for macromolecular structure determination. Acta
3.5 cm-diameter wells, were transfected with 2 ␮g of Qiagen-puri- Crystallogr. D54, 905–921.
fied plasmids with the use of Fugene-6 reagent (Roche). Forty-eight Camus, D., and Hadley, T.J. (1985). A Plasmodium falciparum anti-
hours after transfection, one coverslip was removed from each well gen that binds to host erythrocytes and merozoites. Science 230,
for immunofluorescence assays to assess expression and control 553–556.
of transfection efficiency. Cells were fixed with 2% formaldehyde
CCP4 (Collaborative Computational Project, Number 4)(1994). The
in phosphate buffered saline, and immunoflourescence observed CCP4 suite: programs for protein crystallography. Acta Crystallogr.
by standard methods (Sim et al., 1990). Rosetting assays were pre-
D Biol. Crystallogr. 50, 760–763.
formed on the remaining coverslips in the 3.5 cm-diameter wells.
Two hundred microliters of 10% hematocrit erythrocytes in media Chitnis, C.E., and Miller, L.H. (1994). Identification of the erythrocyte
were added to each well, mixed, and incubated for 2 hr at 37°C. binding domains of Plasmodium vivax and Plasmodium knowlesi
The COS cells were washed three times with phosphate buffered proteins involved in erythrocyte invasion. J. Exp. Med. 180, 497–
saline to remove nonadherent cells prior to visualization and 506.
scoring. de la Fortelle, E., and Bricogne, G. (1997). Maximum-likelihood
heavy-atom parameter refinement for multiple isomorphous re-
placement and multiwavelength anomalous diffraction methods.
Figures
Methods Enzymol. 276, 472–494.
Figures 2A, 2B, 3A, 3C–3E, and 4 were prepared with MolScript
(Kraulis, 1991), Raster 3D (Bacon and Anderson, 1988; Merritt and Duraisingh, M.T., Maier, A.G., Triglia, T., and Cowman, A.F. (2003a).
Murphy, 1994) and POVScript+ (Fenn et al., 2003). Figure 3B was Erythrocyte-binding antigen 175 mediates invasion in Plasmodium
prepared with GRASP (Nicholls et al., 1991) and POVScript+. falciparum utilizing sialic acid-dependent and -independent path-
ways. Proc. Natl. Acad. Sci. USA 100, 4796–4801.
Duraisingh, M.T., Triglia, T., Ralph, S.A., Rayner, J.C., Barnwell, J.W.,
McFadden, G.I., and Cowman, A.F. (2003b). Phenotypic variation of
Supplemental Data
Plasmodium falciparum merozoite proteins directs receptor tar-
Supplemental Data include four figures and four tables and can be
geting for invasion of human erythrocytes. EMBO J. 22, 1047–1057.
found with this article online at http://www.cell.com/cgi/content/
full/122/2/183/DC1/. Duraisingh, M.T., Voss, T.S., Marty, A.J., Duffy, M.F., Good, R.T.,
Thompson, J.K., Freitas-Junior, L.H., Scherf, A., Crabb, B.S., and
Cowman, A.F. (2005). Heterochromatin silencing and locus reposi-
Acknowledgments tioning linked to regulation of virulence genes in Plasmodium falci-
parum. Cell 121, 13–24.
Recombinant RII was produced by EntreMed, Inc. We thank Hong
Fenn, T.D., Ringe, D., and Petsko, G.A. (2003). POVScript+: a pro-
Liang, Tom Chen, and the Molecular Biology and Pharmaceutical
gram for model and data visualization using persistence of vision
Science teams for the production of purified recombinant RII, Mar-
ray-tracing. J. Appl. Crystallogr. 36, 944–947.
garet Luk-Paszyc and Caroline Kisker for assistance with analytical
ultracentrifugation, Reinhard Brossmer for reagents and advice, Freitas-Junior, L.H., Hernandez-Rivas, R., Ralph, S.A., Montiel-
and Michael Becker (beamline X25) and Annie Heroux (beamline Condado, D., Ruvalcaba-Salazar, O.K., Rojas-Meza, A.P., Mancio-
X26C) for support with data collection at the National Synchrotron Silva, L., Leal-Silvestre, R.J., Gontijo, A.M., Shorte, S., and Scherf,
Light Source (NSLS) at Brookhaven National Laboratory. Special A. (2005). Telomeric heterochromatin propagation and histone acet-
thanks to Stephen A. Johnston. The NSLS is supported by the US ylation control mutually exclusive expression of antigenic variation
Department of Energy, Division of Materials Sciences and Division genes in malaria parasites. Cell 121, 25–36.
of Chemical Sciences. This work was supported by an SBIR grant Gilberger, T.W., Thompson, J.K., Reed, M.B., Good, R.T., and Cow-
1R43 AI46209 (to B.K.L.S.), a Leslie C. Quick Jr. Predoctoral Fellow- man, A.F. (2003a). The cytoplasmic domain of the Plasmodium fal-
ship to N.H.T., and the Watson School of Biological Sciences and ciparum ligand EBA-175 is essential for invasion but not protein
the Louis Morin Charitable Trust (L.J.). trafficking. J. Cell Biol. 162, 317–327.
Cell
192

Gilberger, T.W., Thompson, J.K., Triglia, T., Good, R.T., Duraisingh, BAEBL) involved in erythrocyte receptor binding. Mol. Biochem.
M.T., and Cowman, A.F. (2003b). A novel erythrocyte binding anti- Parasitol. 119, 159–168.
gen-175 paralogue from Plasmodium falciparum defines a new Navaza, J., and Saludjian, P. (1997). AMoRe: an automated molecu-
trypsin-resistant receptor on human erythrocytes. J. Biol. Chem. lar replacement program package. Methods Enzymol. 276, 581–
278, 14480–14486. 594.
Gowda, D.C., and Davidson, E.A. (1999). Protein glycosylation in Nicholls, A., Sharp, K.A., and Honig, B. (1991). Protein folding and
the malaria parasite. Parasitol. Today 15, 147–152. association: insights from the interfacial and thermodynamic prop-
Guzman, F., Jaramillo, K., Salazar, L.M., Torres, A., Rivera, A., and erties of hydrocarbons. Proteins 11, 281–296.
Patarroyo, M.E. (2002). 1H-NMR structures of the Plasmodium fal- Orlandi, P.A., Klotz, F.W., and Haynes, J.D. (1992). A malaria inva-
ciparum 1758 erythrocyte binding peptide analogues and protec- sion receptor, the 175-kilodalton erythrocyte binding antigen of
tion against malaria. Life Sci. 71, 2773–2785. Plasmodium falciparum recognizes the terminal Neu5Ac(alpha
Holm, L., and Sander, C. (1993). Protein structure comparison by 2–3)Gal- sequences of glycophorin A. J. Cell Biol. 116, 901–909.
alignment of distance matrices. J. Mol. Biol. 233, 123–138. Rayner, J.C., Galinski, M.R., Ingravallo, P., and Barnwell, J.W.
Jones, T.A., and Kjeldgaard, M. (1997). Electron-density map inter- (2000). Two Plasmodium falciparum genes express merozoite pro-
pretation. Methods Enzymol. 277, 173–208. teins that are related to Plasmodium vivax and Plasmodium yoelii
adhesive proteins involved in host cell selection and invasion. Proc.
Jones, T.R., Narum, D.L., Gozalo, A.S., Aguiar, J., Fuhrmann, S.R.,
Natl. Acad. Sci. USA 97, 9648–9653.
Liang, H., Haynes, J.D., Moch, J.K., Lucas, C., Luu, T., et al. (2001).
Protection of Aotus monkeys by Plasmodium falciparum EBA-175 Rayner, J.C., Vargas-Serrato, E., Huber, C.S., Galinski, M.R., and
region II DNA prime-protein boost immunization regimen. J. Infect. Barnwell, J.W. (2001). A Plasmodium falciparum homologue of
Dis. 183, 303–312. Plasmodium vivax reticulocyte binding protein (PvRBP1) defines a
trypsin-resistant erythrocyte invasion pathway. J. Exp. Med. 194,
Kaneko, O., Mu, J., Tsuboi, T., Su, X., and Torii, M. (2002). Gene
1571–1581.
structure and expression of a Plasmodium falciparum 220-kDa pro-
tein homologous to the Plasmodium vivax reticulocyte binding pro- Reed, M.B., Caruana, S.R., Batchelor, A.H., Thompson, J.K., Crabb,
teins. Mol. Biochem. Parasitol. 121, 275–278. B.S., and Cowman, A.F. (2000). Targeted disruption of an erythro-
cyte binding antigen in Plasmodium falciparum is associated with
Kleywegt, G.J. (1999). Experimental assessment of differences be-
a switch toward a sialic acid-independent pathway of invasion.
tween related protein crystal structures. Acta Crystallogr. D Biol.
Proc. Natl. Acad. Sci. USA 97, 7509–7514.
Crystallogr. 55, 1878–1884.
Rodriguez, L.E., Urquiza, M., Ocampo, M., Suarez, J., Curtidor, H.,
Klotz, F.W., Orlandi, P.A., Reuter, G., Cohen, S.J., Haynes, J.D.,
Guzman, F., Vargas, L.E., Trivinos, M., Rosas, M., and Patarroyo,
Schauer, R., Howard, R.J., Palese, P., and Miller, L.H. (1992). Bind-
M.E. (2000). Plasmodium falciparum EBA-175 kDa protein peptides
ing of Plasmodium falciparum 175-kilodalton erythrocyte binding
which bind to human red blood cells. Parasitology 120, 225–235.
antigen and invasion of murine erythrocytes requires N-acetylneu-
raminic acid but not its O-acetylated form. Mol. Biochem. Parasitol. Sheldrick, G.M., Dauter, Z., Wilson, K.S., Hope, H., and Sieker, L.C.
51, 49–54. (1993). The application of direct methods and Patterson interpreta-
tion to high-resolution native protein data. Acta Crystallogr. D49,
Kraulis, P.J. (1991). MOLSCRIPT: a program to produce both de-
18–23.
tailed and schematic plots of protein structures. J. Appl. Crystal-
logr. 24, 946–950. Sim, B.K., Orlandi, P.A., Haynes, J.D., Klotz, F.W., Carter, J.M., Ca-
mus, D., Zegans, M.E., and Chulay, J.D. (1990). Primary structure of
Liang, H., and Sim, B.K. (1997). Conservation of structure and func-
the 175K Plasmodium falciparum erythrocyte binding antigen and
tion of the erythrocyte-binding domain of Plasmodium falciparum
identification of a peptide which elicits antibodies that inhibit ma-
EBA-175. Mol. Biochem. Parasitol. 84, 241–245.
laria merozoite invasion. J. Cell Biol. 111, 1877–1884.
Lo Conte, L., Chothia, C., and Janin, J. (1999). The atomic structure
Sim, B.K., Chitnis, C.E., Wasniowska, K., Hadley, T.J., and Miller,
of protein-protein recognition sites. J. Mol. Biol. 285, 2177–2198.
L.H. (1994). Receptor and ligand domains for invasion of erythro-
Lovell, S.C., Davis, I.W., Arendall, W.B., 3rd, de Bakker, P.I., Word, cytes by Plasmodium falciparum. Science 264, 1941–1944.
J.M., Prisant, M.G., Richardson, J.S., and Richardson, D.C. (2003).
Sim, B.K., Narum, D.L., Liang, H., Fuhrmann, S.R., Obaldia, N., 3rd,
Structure validation by Calpha geometry: phi, psi and Cbeta devia-
Gramzinski, R., Aguiar, J., Haynes, J.D., Moch, J.K., and Hoffman,
tion. Proteins 50, 437–450.
S.L. (2001). Induction of biologically active antibodies in mice, rab-
MacKenzie, K.R., Prestegard, J.H., and Engelman, D.M. (1997). A bits, and monkeys by Plasmodium falciparum EBA-175 region II
transmembrane helix dimer: structure and implications. Science DNA vaccine. Mol. Med. 7, 247–254.
276, 131–133.
Smith, J.D., Chitnis, C.E., Craig, A.G., Roberts, D.J., Hudson-Taylor,
Marchesi, V.T., Tillack, T.W., Jackson, R.L., Segrest, J.P., and Scott, D.E., Peterson, D.S., Pinches, R., Newbold, C.I., and Miller, L.H.
R.E. (1972). Chemical characterization and surface orientation of (1995). Switches in expression of Plasmodium falciparum var
the major glycoprotein of the human erythrocyte membrane. Proc. genes correlate with changes in antigenic and cytoadherent pheno-
Natl. Acad. Sci. USA 69, 1445–1449. types of infected erythrocytes. Cell 82, 101–110.
Mayer, D.C., Kaneko, O., Hudson-Taylor, D.E., Reid, M.E., and Snow, R.W., Craig, M., Deichmann, U., and Marsh, K. (1999). Esti-
Miller, L.H. (2001). Characterization of a Plasmodium falciparum mating mortality, morbidity and disability due to malaria among
erythrocyte-binding protein paralogous to EBA-175. Proc. Natl. Africa’s non-pregnant population. Bull. World Health Organ. 77,
Acad. Sci. USA 98, 5222–5227. 624–640.
Mayor, A., Bir, N., Sawhney, R., Singh, S., Pattnaik, P., Singh, S.K., Su, X.Z., Heatwole, V.M., Wertheimer, S.P., Guinet, F., Herrfeldt, J.A.,
Sharma, A., and Chitnis, C.E. (2004). Receptor-binding residues lie Peterson, D.S., Ravetch, J.A., and Wellems, T.E. (1995). The large
in central regions of Duffy-binding-like domains involved in red cell diverse gene family var encodes proteins involved in cytoadher-
invasion and cytoadherence by malaria parasites. Blood 105, ence and antigenic variation of Plasmodium falciparum-infected
2557–2563. erythrocytes. Cell 82, 89–100.
Merritt, E.A., and Murphy, M.E.P. (1994). Raster3D version 2.0—a Taylor, H.M., Grainger, M., and Holder, A.A. (2002). Variation in the
program for photorealistic molecular graphics. Acta Crystallogr. expression of a Plasmodium falciparum protein family implicated
D50, 869–873. in erythrocyte invasion. Infect. Immun. 70, 5779–5789.
Miller, L.H., Baruch, D.I., Marsh, K., and Doumbo, O.K. (2002). The Terwillliger, T.C. (2000). Maximun likelihood density modification.
pathogenic basis of malaria. Nature 415, 673–679. Acta Crystallogr. D56, 965–972.
Narum, D.L., Fuhrmann, S.R., Luu, T., and Sim, B.K. (2002). A novel Thompson, J.K., Triglia, T., Reed, M.B., and Cowman, A.F. (2001). A
Plasmodium falciparum erythrocyte binding protein-2 (EBP2/ novel ligand from Plasmodium falciparum that binds to a sialic
Crystal Structure of a Malaria Vaccine Candidate
193

acid-containing receptor on the surface of human erythrocytes.


Mol. Microbiol. 41, 47–58.
Triglia, T., Thompson, J.K., and Cowman, A.F. (2001). An EBA175
homologue which is transcribed but not translated in erythrocytic
stages of Plasmodium falciparum. Mol. Biochem. Parasitol. 116,
55–63.
Triglia, T., Duraisingh, M.T., Good, R.T., and Cowman, A.F. (2005).
Reticulocyte-binding protein homologue 1 is required for sialic
acid-dependent invasion into human erythrocytes by Plasmodium
falciparum. Mol. Microbiol. 55, 162–174.
Vassilev, L.T., Vu, B.T., Graves, B., Carvajal, D., Podlaski, F., Fili-
povic, Z., Kong, N., Kammlott, U., Lukacs, C., Klein, C., et al. (2004).
In vivo activation of the p53 pathway by small-molecule antago-
nists of MDM2. Science 303, 844–848.
Weeks, C.M., and Miller, R. (1999). The design and implementation
of SnB version 2.0. J. Appl. Crystallogr. 32, 120–124.
Word, J.M., Lovell, S.C., LaBean, T.H., Taylor, H.C., Zalis, M.E., Presley,
B.K., Richardson, J.S., and Richardson, D.C. (1999). Visualizing and
quantifying molecular goodness-of-fit: small-probe contact dots
with explicit hydrogen atoms. J. Mol. Biol. 285, 1711–1733.

Accession Numbers

The atomic coordinates and structure factors have been deposited


in the Protein Data Bank (PDB entry codes 1ZRL and 1ZRO).

You might also like