You are on page 1of 32

The Organism Prearranged Recognition Theory

By: Luz P. Blanco Ph.D.

Light White Innovation Technology, Ann Arbor, MI.

E-mail: lwinntech@gmail.com.

Phone: (734) 485-4907

Summary

Presented herein is a theory to explain how organisms' interactions are


prearranged. This theory provides the molecular basis for the initial event in the
establishment of these interactions and should aid in the development of better
vaccine strategies and new, efficient therapies against infectious diseases and
cancer. The theory is as follows: “The recognition and interaction that occur
between organisms is prearranged. There are soluble components derived
from one organism that recognize and allow interaction only with those
unique organisms that have the specific receptors for those soluble
components”. The organisms’ prearranged recognition theory (OPRT) can be
specifically applied to host-microbe interactions where most microbes are coated
(opsonised) by soluble components (opsonins) from the host, but there are also
some microbes that can bypass the host opsonization by directly expressing
receptors for the host cells, thus directly mimicking the host opsonins. The
receptors involved in an organism's interactions, their specificity and repertoire
are based mainly, but not exclusively, on the epitopes of the saccharides located
in the glycoproteins, glycolipids and polysaccharides (glycans) that are highly
abundant on the surfaces and extracellular components of living organisms. This
prearrangement may provide an explanation of species-specific interactions and
several other important and previously unresolved phenomena, such as
hyperinfectivity, tissue tropism, higher sensitivity of people with type O-blood to
certain diseases in comparison with other blood groups, and tumoral cell
promiscuity. The lipid raft domain in the cellular membrane is the main location
where interactions will trigger cellular responses. I herein propose some
alternative routes to modify an organism's interactions as well as the possible
consequences of these manipulations. It is a novel theory regarding the degree
to which an organism’s interactions are prearranged and the importance of
saccharide epitopes in the molecular mechanism of this prearrangement.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
Organisms’ prearranged recognition theory (OPRT):

“The recognition and interaction that occur between organisms is


prearranged. There are soluble components (opsonins) derived from one
organism that recognize and allow interaction only with those unique
organisms that have the specific receptors for those opsonins”.

Based on this theory, there are two types of organisms or microbes for a
specific opsonising host-organism: those for which interactions could not be
established (type I) and those for which interactions could be productively
established (type II). Type I organism which is unable to establish interaction,
will thus go unnoticed because it is without host-specific opsonin receptors and is
unable to directly express receptors specific for the host. Type I organisms for a
human host-organism are those multiple microbes that fail to establish interaction
with us but they will interact (invade, infect, colonize, adhere, etc.) with other
species. Type II organisms, which will productively establish interactions with a
given host-organism are either opsonized by the host opsonins or are able to
directly express receptors for the specific host cells (Figure 1). Depending on
their ability to disseminate and overgrow inside the host, type II organisms can
either be beneficial or deleterious. The latter affect the energy homeostasis of
the host cells and tissues and produce damage and eventually disease. Factors
such as the nutritional or the immune state of the host will determine if a
beneficial organism can be transformed into a deleterious organism. The
damage-response framework theory explains the possible outcomes, once the
initial interactions are established (Casadevall and Pirofski 2003). Antibodies—
and in particular, natural antibodies—play an instrumental role as the main
opsonins in the organisms’ interactions, but so also do other abundant and
ubiquitous glycoproteins and glycolipids. The interactions that happen between
organisms and the immune system follow this theory; however, the type of
receptors and antigens involved in these types of interactions are mainly—but
not exclusively—proteins and peptides. In cell membranes, the lipid raft
platforms are the location where the first events involved in recognition and
interactions are inititated including the vigilance that triggers signalling cascades.

Molecular basis for the repertoire and diversity of receptors:


The diversity and repertoire specificity of opsonin substances, as well as the
organisms’ receptors, are based mainly on the saccharide epitopes present in
the glycans, glycoproteins, glycolipids, sugar coats, capsules, cell walls, etc. The
importance of saccharides in the interaction between different organisms and
elements can be deduced from the mortality rates and high severity of diseases

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
related to genetic disorders in the saccharides' biosynthetic pathways (De
Praeter, Gerwig et al. 2000; Freeze 2001; Freeze 2002; Freeze 2006). In the
mucosal system, glycoproteins, such as antibodies and mucins, are the main
prearranged recognition opsonins. In serum, lymphatic fluid, and tissues, the
opsonins can be antibodies, as well other glycoproteins and glycolipids. The
envelope antigens containing saccharide epitopes that are the receptor targets
for opsonization in organisms such as bacteria, fungi, parasites, and viruses are
commonly used in serology to classify their serotype group. The serology of
organisms is related to the surface antigens (usually saccharides) expressed,
such as the lipopolyssacharide’s (LPS) O antigen in Gram-negative bacteria and
this characteristic also determines their ability to productively interact with a
specific host. These receptors are what host opsonins can detect or recognize to
allow interactions with a foreign organism. In fact, applying the OPRT, we can
understand why serotypes are useful in classifying the host's susceptibility range
or the organism's species-specificity. Antibodies (which are glycoproteins) could
basically be considered as the main tools for an organism's interaction because
they allow the interaction of foreign antigens and microorganisms with cells.
Supporting the central role of glycoconjugates in human gut interactions,
saccharides resemble microbial receptors, bacterial LPS, and the structure of
tumoral antigens (Henry 2001), thus ensuring that these structures are familiar to
the host. Human gut mucins, which are components of the mucus, and
glycoproteins are highly diverse and provide a broad repertoire of host receptors
(Robbe, Capon et al. 2004). The mucus layer also has been characterized as
instrumental to the structure and stability of the normal gut flora (Sonnenburg,
Angenent et al. 2004).

Carbohydrates provide the molecular basis for specific cell-cell


recognition:
In the model of species-specific cell-cell recognition of the marine sponge, the
role of carbohydrate interaction (glycan-glycan) has been clearly established
(Bucior and Burger 2004; Bucior and Burger 2004; Bucior, Scheuring et al.
2004). The glycans allow species-specific reconstitution of the sponge structure
from individual disaggregated cells. Importantly, the strength of the carbohydrate-
carbohydrate interaction is as strong as a protein-protein interaction, as was
determined using atomic force microscopy (Popescu, Checiu et al. 2003; Bucior,
Scheuring et al. 2004). Carbohydrates have a self-recognition capability, and the
specificity of these interactions has been demonstrated using glycans derived
from marine sponges, a surface plasmon resonance detection system, and
atomic force microscopy (Gradimir 1999; Haseley, Vermeer et al. 2001; Misevic,
Guerardel et al. 2004). In a multicellular organism and a membrane bilayer
environment, protein-glycolipid interactions require a lipid bilayer environment for
proper presentation of the saccharide epitopes with higher affinity and specificity
to proteins (Evans and MacKenzie 1999).

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
Carbohydrates located in the surface of cells, such as glycoproteins,
gycolipids, gangliosides and others, are already known to participate in cell-cell
adhesion and cell response in highly diverse events (Brewer, Miceli et al. 2002;
Hakomori 2004; Hakomori 2004b; Eklund and Freeze 2005). A number of
glycoproteins participate in interactions required for a proper immune system
response (Tsuboi and Fukuda 1998; Tsuboi and Fukuda 2001; Rudd, Wormald
et al. 2004; Arnold, Wormald et al. 2007). Moreover, glycosylation in diverse
viruses plays a role in processes such as attachment, infectivity, entry, and
replication, which also require host glycoproteins (Vigerust and Shepherd 2007).
Some representative examples of microbe- or toxin-host interactions where
saccharides play a role in attachment or infection are depicted in Table 1.
In humans and mice, muscle integrity depends on glycosylation because
mutations in glycosyltransferases have been associated with some muscular
dystrophy cases (Martin and Freeze 2003). Also, endocrine system alterations
are due to congenital glycosylation defects (Miller and Freeze 2003). Overall,
there is strong support for the role of saccharides as the host’s receptors for the
interactions to proceed and for the specificity of interactions even within the
tissues and organs of the host itself. The molecular mechanism involved in
particular in the host-pathogen interactions described here probably has long
been hindered because 2-3% of the total genome is involved in saccharide
biosynthesis pathways that are not yet well characterized (Freeze 2001). Also,
the genetic manifestations of saccharide disorders that affect specific host-
pathogen interactions probably require multiple simultaneous gene changes that
are not necessarily linked.

Molecular basis for species-specificity, tropism, and hyperinfectivity


phenomenon:
The organism prearranged recognition theory gives the molecular basis for
species-specificity of colonization, infection phenomenon, and also for in vivo cell
tropism of some infectious or deleterious organisms. Species-specificity is given
by an organism’s ability to express appropriate receptors for the host's opsonins,
and this allows the host to interact specifically with only those organisms. Also,
the cell tropism of certain organisms is determined by recognition of the
opsonins, for which there are host tissues or cell-specific receptor(s) that allow
an interaction to proceed only with those particular tissues. Applying the OPRT
to understand the hyperinfection phenomenon is straightforward, where
pathogens that have been inside a host acquire a more virulent phenotype and
already carry the host’s opsonins that optimizes them for interaction with any
subsequently appropriate hosts. The enteropathogens Vibrio cholerae (Merrell,
Butler et al. 2002) and Citrobacter rodentium (Wiles, Dougan et al. 2005; Bishop,
Wiles et al. 2007) display hyperinfection, or host-adapted phenotypes. The
hyperinfectious phenotype is acquired after host passage and is temporarily
acquired because it is gradually lost once the pathogen is outside the host. The
host’s shield of opsonins probably aids the organism in resisting stressing

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
conditions inside and outside the host, such as the acid pH, the temperature, the
osmolarity, and the proteases. Opsonins also may help organisms to build
biofilms, where they are in a more protected state, compared with their
planktonic, free-living state. In support of this hypothesis, it has been recently
shown that aggregates or biofilm arrangements of Vibrio cholerae contribute both
to increased infectious capability and greater environmental persistence, when
compared to free-swimming planktonic cells (Faruque, Biswas et al. 2006).
The OPRT explains why it is usually difficult to reproduce the species-
specificity property of interactions in vitro, probably because species-specific
opsonins are not present (the fetal serum used in culture is from a different
species), changing the organism's coat of opsonins and rendering them unable to
interact optimally or—the opposite—making them artificially able to interact with a
non-host organism. For example, the in vitro ability of different Salmonella
serotypes to resist macrophages does not correlate with their in vivo virulence
(Watson, Paulin et al. 2000). However, there are several non-invading bacteria
in vivo that are invasive to some cells like tumor cells in vitro.

Molecular basis for promiscuity of tumoral cells:


The more promiscuous an organism is (in other words, the more interactions it is
able to establish), the more chances it has to develop a deleterious interaction or
to acquire pathogenic organisms. The promiscuity is a function of the repertoire
or the diversity of the opsonins and receptors each organism expresses and
carries. Diversity can probably be induced, and it is different between even
individuals of the same species. Tumoral cells are one of the most promiscuous
of cell types because they change their "coat" or outer shell, unlike normal, non-
transformed cells; thus, tumoral cells express a variety of saccharides on their
surfaces, presenting a broader repertoire of receptors and opsonins, such as
novel membrane-associated mucins (Walsh, Young et al. 2007). These novel
envelope glycoproteins allow tumoral cells to be colonized by different types of
bacteria and viruses, and to interact with each other and with other normal cells
both in vivo and in vitro with high avidity and efficiency (Figure 1). Tumoral cells
express organisms’ interaction molecules in their membranes, so they are
prompted to adhere and to interact, and they express mucin and mucin-like
glycoproteins containing N- and O-glycans that play a role in cell transformation,
cell adhesion, invasion, immune escape, and metastasis (Brockhausen 2003;
Hollingsworth and Swanson 2004). Understanding the chemistry of carbohydrate
biosynthesis in cancer cells should help in defining innovative therapies against
them (Paulsen and Brockhausen 2001). Other glycoproteins, particularly, those
with a high sialic acid content, are also associated with the inherent properties of
cancerous cells (Dall'Olio and Chiricolo 2001; Ono and Hakomori 2003; Suzuki,
Suzuki et al. 2005), and these properties have recently been used to target
cationic polymers against certain mucosal tumoral cells (Azab, Kleinstern et al.
2008).

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
Vigilance toward and detection of type II organisms:
The vigilance toward and detection of type II organisms by the host are mediated
through receptors able to recognize saccharide epitopes—for example,
gangliosides and cell differentiation molecules or lectin-like molecules in the
cellular membrane of innate immune cells. Also, toll-like receptors that recognize
microbe-derived molecules play a role in host vigilance, as is discussed below. I
have deduced from my study with Vibrio cholerae, which is a human species-
specific pathogen (Blanco and DiRita 2006; Blanco and DiRita 2006a), that
gangliosides are host receptors playing a role in the organism's interaction. The
main virulence factor of this pathogen is the cholera toxin, the host receptor of
which is the ubiquitous glycosphingolipid ganglioside GM1. The ganglioside
GM1 has the same structure in different species, but only humans are sensitive
to V. cholerae infection—and especially subjects having blood type O (Barua and
Paguio 1977). Therefore, I propose that a cholera toxin-GM1 link exists in the
human intestine that is missing in other species but is present especially in the
more sensitive blood type O individuals, allowing V. cholerae to colonize and to
produce the cholera disease. Previously it has been demonstrated that the type
O blood group antigen in host cells is unable to directly interact with cholera toxin
better than A or B blood group antigens (Sugii 1990) can. However, nonimmune
human antibodies (sIgA or IgG), which are glycoproteins (Arnold, Wormald et al.
2007) from healthy non-immunized people, recognize V. cholerae and also
interact with GM1 (Blanco and DiRita 2006a). It is possible therefore to envision
these antibodies as host opsonins which bridge the microbe with the host cells.
Most humans at 6 months of age have cross-reactive anti-GM1 antibodies in
their serum. The generation of these antibodies coincides with the appearance
of anti-blood group antigen and anti-Forssman antibodies; this has been related
to immune response development against microbes or, applying the OPRT,
probably the development of the organism's interactions (Alaniz, Lardone et al.
2004). Recently, analyses of antibodies against tumoral antigens (mostly
saccharide epitopes) demonstrate that they are only natural antibodies or germ-
line coded IgM (Vollmers and Brandlein 2006; Vollmers and Brandlein 2007);
strikingly, no affinity-maturated, anti-tumor-specific IgG antibodies were found.
Also, there are anti-saccharide antibodies (opsonins) present in the gut
environment about which “a protective role against pathogens” has been
proposed (Hansen, Pedersen et al. 2005), but they are likely allowing the
colonization or interaction with the highly abundant, normal mucosal flora. In a
different context, antibodies against gangliosides have been associated with
some neurological diseases, but there is no direct correlation between the
presence of the antibodies and the manifestation of disease (Garcia Guijo,
Garcia-Merino et al. 1995); however, the antibodies' affinity, amount, or other
inherent characteristics could be involved in neurological pathology triggering.
The glycosynapse concept developed by Hakomori supports the proposal
that gangliosides play a functional role in organisms’ interactions (Hakomori
2002). In the review cited, Hakomori shows how gangliosides in clusters or

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
microdomains, together with functional membrane proteins, play a significant role
in processes such as cell-adhesion, development, and differentiation and in
tumor progression (Hakomori 2002). The role of gangliosides in pathogen
detection and host-pathogen interactions is suggested by the long list of
pathogens that are able to recognize the carbohydrate portion of the
glycosphingolipids (Schengrund 2003). Importantly, some viruses, toxins, and
bacteria use gangliosides as receptors (Blanco 2008) to infect cells. Also, in
order to adhere, several pulmonary pathogenic bacteria recognize and bind to
glycosphingolipids, and common saccharide moieties in vitro abrogate their
adherence (Krivan, Roberts et al. 1988; Thomas and Brooks 2004; Thomas and
Brooks 2004). Moreover, ganglioside GM1 is able to present antigens (Nashar,
Betteridge et al. 2002) to activate secretion of neurotrophins (Rabin, Bachis et al.
2002) together with other gangliosides to insert into membranes (Schwarzmann
2001; Lauc and Heffer-Lauc 2006), and they also stabilize lipid domains in
cellular membranes (Sonnino, Mauri et al. 2007). Remarkable, some tumoral
cells secrete gangliosides; this ability probably allows them to modify and to
improve their interactions with other cells as well as to regulate the immune
responses (Shen, Falahati et al. 2005).
If there is the ability to sense or to be recognized, then there is the
ability to interact and to produce cellular responses. This property of the
interactions facilitates vigilance against any organism that can trespass,
overgrow, or invade deeper portions of the body that might remain sterile. The
sensing is accomplished through detection of the microbe's derived molecules by
toll-like receptors, gangliosides, or CD receptors, and cellular responses or
signaling events are originated in lipid raft platforms in the membranes. These
kinds of receptors are expressed by innate immune cells that, when they sense
any potential deleterious organism, trigger an alarm response, especially when
the detection or interaction is happening in tissues that must remain sterile and is
disrupting the energy homeostasis balance. For example, in dendritic cells the
CD209 (DC-SIGN) receptor recognizes the N-acetylglucosamine sugar residues
within the core LPS of diverse bacteria (Zhang, Snyder et al. 2006). The alarm
response is based on the type, location, amount of cytokines, and heat shock
proteins (HSP) produced or expressed by the cells and will trigger a specific
defensive immune response—for example, a cellular- and humoral-specific
immune response. Strikingly, the specific immune responses also require
prearrangement to take place. Accordingly, antigens interact with MHC molecule
receptors in order to be presented to lymphocytes that might have the precise
prearranged receptor, and antibodies that are already present recognize specific
antigens. Clearly, prearrangement in a broad variety of interactions is constantly
present.
The interaction of immune-specific cells and antibodies with the triggering
target antigens or the recognized microbe produces a defensive response
towards elimination of the organism because it allows interaction of the invading
organism or antigen with specific immune effector cells; however, low-affinity and

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
high-multiplicity interactions through saccharide moieties in antibodies
(erroneously considered “non-specific” types of interactions) allow the
prearranged interaction to proceed without an aggressive component. This
proposition is clearer to visualize by analyzing sIgA interaction with normal
colonizing flora in the gut. High-affinity specific sIgA is probably involved in the
immune exclusion phenomena of pathogens, but normal flora microbes—which
exist as a highly stable ecosystem—are coated with sIgA. They are not excluded
from the intestinal lumen and neither do they induce a specific immune effector
response (Kroese, de Waard et al. 1996; Mestecky, Russell et al. 1999), despite
the fact that they are processed by mucosal dendritic cells (Macpherson and Uhr
2004). As an example, gut antigen-specific sIgA against Reovirus type 1 Lang is
able to block and to neutralize viral infections (Hutchings, Helander et al. 2004).
However, in a murine model of Streptococcus pneumoniae colonization where
the bacteria interact with the host, antibodies do not participate in bacterial
clearance (McCool and Weiser 2004).

Glycoproteins that play a role in organisms’ interactions:


The role of glycans in mucins for colonization, biofilm formation, and cell-cell
interaction phenomena is well-documented. Mucins are important for interaction
phenomena such as colonization and adhesion of a variety of microorganisms,
not only in the gut but also in the airway apparatus, in the oral micro-biota and
even between tumoral cells (Alice and Molakala 1996). A septic shock in mice
infected with the human species-specific pathogen Salmonella typhi can be
induced only when the bacteria are pre-opsonized with mucin (Sein, Cachicas et
al. 1993). The S. typhi-mucin inoculums are able to survive inside murine
peritoneal macrophages, contrary to nude S. typhi that go unnoticed in mice and
are eradicated without causing disease, even in immune-deficient SCID mice
(Sein, Cachicas et al. 1993). Mucin gives a new coating to the bacterium that
bridges the interaction of S. typhi with murine cells. Importantly, the expression
of certain envelope-mucin molecules has been associated with the malignancy of
tumoral cells and metastasis. In fact, inhibition of MUC-4 (a surface associated
mucin) expression in a highly aggressive pancreatic tumor cell line suppresses
metastasis in mice (Singh, Moniaux et al. 2004), and MUC1 expression
suppressed by RNA interference reduces the growth rate and metastasis ability
of human pancreatic cancer cells (Tsutsumida, Swanson et al. 2006).
Applying the OPRT, it is possible to understand why there are several
pathogens—not only V. cholerae, but Norwalk virus (Hutson, Atmar et al. 2002),
Escherichia coli O157 (Blackwell, Dundas et al. 2002), and Helicobacter pylori
(Mentis, Blackwell et al. 1991) infections—found more frequently in type O blood
group subjects. People of the type O blood group have antibodies against the A
and B saccharide antigens that cross-react with microbial and viral envelope
saccharide epitopes (Springer 1970), giving type O blood group antigen
individuals more opportunities to interact with different organisms or to enhance
interactions with those that are prearranged or express the right receptors. Other

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
examples of antibody- enhancing interactions involve IgG and IgE, which can
have a role in antigen-sampling through the epithelial barriers; particularly critical
is the role of lumenal IgE in people with food hypersensitivity (Berin, Li et al.
2006; Li, Nowak-Wegrzyn et al. 2006) and in pregnancies where non-immune
IgG enhances infection of the placental erythrocytes with Plasmodium
falciparium (Flick, Scholander et al. 2001).
The sIgA is important for the microorganism’s colonization of oral, airway
and gut environments together with mucus. The amount of sIgA in the gut’s
lumen is determined by the type and presence of bacteria (Macpherson,
Hunziker et al. 2001; Jiang, Thurnheer et al. 2004; Haghighi, Gong et al. 2006;
Ohashi, Hiraguchi et al. 2006). Secretory IgA and mucins form high-molecular-
weight aggregates or networks that allow microorganism colonization in the form
of biofilms. Another example: sIgA and mucins promote interaction of
Salmonella typhimurium with rat intestines (Magnusson and Stjernström 1982).
The presence of sIgA in gut biofilms has been characterized in different species
(Palestrant, Holzknecht et al. 2004), but the repertoire of sIgA is limited (Stoel,
Jiang et al. 2005; Stoel, Evenhuis et al. 2008); however, glycans in the sIgA
structure allow the innate recognition of microbe surface saccharides, broadening
the range of interactions sIgA is able to establish (Wold, Mestecky et al. 1990;
Royle, Roos et al. 2003; Perrier, Sprenger et al. 2006). Also, gut sIgA has two
different origins; one is dependent on (75%) and the other is independent (25%)
of T lymphocytes (Kerr 2000). This dichotomy may help in understanding
conflicting data about the protective role of sIgA. But supporting a role of sIgA in
establishment of organism interactions, sIgA has been proposed as an
endogenous adjuvant (a promoter of element interactions) because antigenic
epitopes expressed in the secretory component of the IgA can induce antigen-
specific systemic and mucosal immune responses (Corthesy, Kaufmann et al.
1996), sIgA has immune modulatory effects similar to a common adjuvant
(Favre, Spertini et al. 2005), and because it can enhance Vibrio cholerae
transcytosis through mucosal antigen-sampling, M-like cells (Blanco and DiRita
2006a).
There are so many examples of host glycoproteins that are important for
the adhesion or attachment processes of viruses, bacteria, fungi, and parasites
that they are difficult to fully enumerate. For example, fibronectin (glycoprotein)
and other extracellular matrix components have been described as important
host elements in pathogen-host interactions, and fibronectin is also able to
interact with ganglioside GM1 (Spiegel, Schlessinger et al. 1984), so fibronectin
is a perfect host opsonin or adjuvant molecule because it bridges the organism’s
interactions. Some gut bacteria are able to interact with mucins and even
express mucin-like proteins in their surface. Also, epithelial cells express
receptors for mucin-like proteins which are important for the promotion of
interactions (Karjalainen, Barc et al. 1994). More examples of host glycoproteins
involved in host-pathogen interactions are depicted in Table 2.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
Manipulating organisms’ interactions:
The prediction is that most of the species-specificity interactions are probably
ruled by the OPRT. One important application of the OPRT is that interactions
between organisms can be modulated, changing the saccharides exposed on
surfaces or expressed by opsonin molecules in the host. This could be exploited
to develop more efficient antibiotics and anti-tumoral drugs with a broader
spectrum by creating saccharide synthesis inhibitors and using enzymes involved
in polysaccharide synthesis, or free saccharides as competitors. Conditions that
affect saccharide interactions, such as changing the pH or the ionic strength
environment, should also block the attachment between tumoral cells or
pathogens and the host cells.
Most organism-host interactions are developed in a lipid raft environment
where clusters of receptors and sugar epitopes are produced. This is why, for
the development of preventive or therapeutic saccharide therapies, the "multi-
valent saccharides" approach should be applied (Schengrund 2003). Non-toxic
carriers of multivalent saccharides, such as dendrimers, can be used; however,
they should be designed specifically to each pathogen and their pathogen-neut-
ralizing ability needs to be tested (Schengrund 2003; Anne Imberty 2008). For
example, mouse influenza pneumonitis can be prevented by polyvalent sialic
acid dendrimers against certain virus subtypes (Landers, Cao et al. 2002). A re-
cently described lectin microarray technique may be useful in characterizing the
organisms’ glycomes (Hsu, Pilobello et al. 2006). Lipid raft domains are impor-
tant not only for cell activation but also for cell-cell interaction and the organisms’
interactions. In this GM1 ganglioside-enriched environment, signaling proteins
are organized, and many microbes use lipid rafts as gateways for invasion or in-
teraction with eukaryotic cells (Manes, del Real et al. 2003; Stuart, Webley et al.
2003; Zaas, Duncan et al. 2005; Blanco 2008). Even immune cell interaction
happens through the immunological synapses that are located in lipid raft do-
mains (Monks, Freiberg et al. 1998; Saito and Yokosuka 2006).
An interesting and encouraging finding is that the adherence in vitro of
enteropathogenic bacteria to epithelial cells is abrogated by prebiotic
galactooligosaccharides (Shoaf, Mulvey et al. 2006), and peptides that recognize
hyaluronic acid prevent wound infections by staphylococcal bacteria in a mouse
model of infection (Zaleski, Kolodka et al. 2006). Using a sialyltransferase
inhibitor in a mouse model, the pulmonary metastasis of a mouse colon
adenocarcinoma was blocked, probably through inhibition of the tumor cells-
platelets interaction because of a reduction in tumor sialic acid residues (Kijima-
Suda, Miyamoto et al. 1986; Kijima-Suda, Miyazawa et al. 1988). Another sialic
acid incorporation inhibitor also was able to inhibit hepatic metastases in a
mouse model (Wagner, Thomas et al. 1990). And recently heptyl α-D-mannose
–high affinity ligand of the FimH type 1 pili from uropathogenic Escherichia coli-
has been shown to inhibit biofilm formation, bacterial adhesion and invasion of
host cells (Wellens, Garofalo et al. 2008). These data encourage the field of

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
prophylactic and therapeutic drug development based on saccharide moieties
modification or recognition.
However, a warning: we must be very careful manipulating saccharide
interactions and always keep in mind that they will determine the ability of
organisms and cells to interact, so, in changing saccharides, we can create new
repertoires of possible interactions and also change species-specificities, create
organisms that can be highly deleterious for different hosts, or interfere with host
immune cell-cell interactions. In fact, the glycosylation levels in the organism,
per se, are critical for interactions to proceed. For example, in elderly humans
augmented glycosylation of critical proteins may cause poor activation of their
immune cells. T cells from aging mice also show defects in activation correlated
with hyperglycosylation patterns; once the modified glycoproteins are
enzymatically corrected, T cell activation is restored (Sadighi Akha and Miller
2005; Sadighi Akha, Berger et al. 2006). Also, the humoral immune response is
impaired in transgenic mice that overexpress O-linked glycans on T cell
membranes because the T-B cell interaction is affected (Tsuboi and Fukuda
1998). Even at the level of transcription regulation in the T and B lymphocyte
activation, the O-glycosylation of NFAT and NFκB is required for transcription
factor translocation into the nuclei and for the induction of IL-2 cytokine
production, as was recently described (Golks, Tran et al. 2007). On the other
hand, excessive deglycosylation can also be deleterious—in fact, deglycosylation
of serum IgA seems to be related to IgA nephropathy, the enhanced attachment
of IgA to mesangial cells, and the induction of apoptosis and proliferation
reduction in these cells (Amore, Cirina et al. 2001; Gao, Xu et al. 2007).
Moreover, in autoimmunity process development, saccharide epitope recognition
may have an important role (Buzás, György et al. 2006). So it is clear that the
state of glycosylation is delicately balanced in nature, and is not straightforward
to manipulate.
Therapies with saccharides also may have a detrimental effect, interfering
with required immune system interactions of innate or adaptive mechanisms,
such as the undesired effect that is produced by glucuronoxylomannan-derived
saccharide from the capsule of the Cryptococcus neoformans fungus. This
saccharide is produced during C. neoformans infections and interferes with
surfactant protein D recognition and agglutination of acapsular opportunistic
pathogens, enhancing the virulence potential of the fungus and probably of other
pathogens in vivo (van de Wetering, Coenjaerts et al. 2004). In fact,
Pseudomonas aeruginosa is opsonized by the surfactant protein D, and this
enhances its phagocytosis and destruction by alveolar macrophages (Restrepo,
Dong et al. 1999). Another example: in vitro studies have demonstrated that
saccharide residues in the C1 inhibitor that recognize LPS can prevent septic
shock. This interaction inhibits up-regulation of TNFα production (Liu, Cramer et
al. 2005), so exogenously added saccharides in this particular case can interfere
with an inherent protective pathway, and that could lead to septic shock.
Nonetheless, certain polysaccharides have robust immune-modulator properties

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
and can be used as therapeutic agents (Cisneros, Gibson et al. 1996; Tzianabos
2000; Balachandran, Pugh et al. 2006). The overall conclusion is that
glycosylation is a clue to interactions between organisms, and great caution is
advisable in its manipulation.
Overall in this work, a novel hypothesis is presented regarding both the
degree to which organisms’ interactions are prearranged and the importance of
saccharide epitopes in the molecular mechanism of this prearrangement.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
Acknowledgments
I would like to emphasize that this theory is the result of my experiences as a
scientist and that I have created it thanks to the constant support and help of my
husband. Thus, I would like to dedicate this theory to my husband, Victor A.
Aguero. I would also like to thank all the people who have touched my life: thank
you for being there for me because each of you has inspired me in some way
and my special thanks to Pat Gold and Shraddha Nigavekar for their insightful
editorial contributions and support.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
14

Figure 1: Diagram depicting the organisms’ prearranged recognition


theory (OPRT) in a simplified view. Type I organisms are not opsonized nor
are they able to express a host-specific receptor; thus, type I organisms are not
able to interact with that specific host-organism and they will go unnoticed.
Examples of type I organisms for the human host-organism are the microbes
that affect species other than humans. Host-specific opsonins recognize type II
organisms or type II organisms directly express host-specific receptors. These
type II organisms’ characteristics allow interaction with that specific host.
Tumoral host cells express additional envelope receptors mimicking the host
opsonins and host-specific receptors, which molecularly explains their higher
promiscuity or higher ability to interact with each other and normal cells better
than normal cells.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
15

Table 1: Saccharides involved in microbe- or toxin-host interactions.

Microbe or toxin/tissue Process or molecules involved Reference


cell interaction
Acanthamoeba keratitis/ A mannose-binding lectin from (Yang, Cao et al.
corneal epithelial cells the parasite is involved in 1997; Cao,
interaction with host cells Jefferson et al.
1998; Garate,
Marchant et al.
2006)
Actinomyces naeslundii/ Saccharides in gangliosides (Brennan, Cisar et
colonizing oral epithelium are recognized by the al. 1984; Brennan,
pathogen Joralmon et al.
1987)
Arcanobacterium Bacterial neuraminidase (Jost, Songer et al.
pyogenes/ adhesion to involved in host cell adhesion 2002)
host cells
Bacteria (Yersinia and SFR-3 expressed in (Cipollo, Awad et al.
Mycobacterium)/ infection nematodes' secretory cells 2004; Höflich,
of nematodes plays a role in surface Berninsone et al.
glycoconjugates composition 2004)
Burkholderia cepacia/ Dextran and xylitol inhibit (Sajjan, Moreira et
human airway explants adherence to human airway al. 2004)
transplant
Candida albicans/ model Fungal glucanase adhesin (Sandini, La Valle et
of haematogenously has a role in hyphal al. 2007)
disseminated candidiasis morphogenesis, adherence to
in mice plastic, and pathogenesis
Candida albicans/ host Fungal O- and N-glycosylation (Munro, Bates et al.
epithelial and innate required for interactions with 2005; Bates,
immune cells mannose-binding receptor Hughes et al. 2006;
and TLR Netea, Gow et al.
2006)
Candida glabrata/ human Fungal lectin involved in (Cormack, Ghori et
epithelial cells adherence al. 1999)
Cry toxin from Bacillus Saccharide biosynthesis of (Griffitts, Huffman et
thuringiensis/ intoxication pathways related with al. 2003; Griffitts,
of target nematode cells gangliosides in nematodes Haslam et al. 2005;
are required for toxin Barrows, Griffitts et
sensitivity al. 2006)
Filovirus/ entry into Entry into human (Takada, Fujioka et
human macrophages macrophages is promoted by al. 2004)
C-type lectin specific for
galactose and N-
acetylgalactosamine
Helicobacter pylori/ Saccharides expressed in (Gustafsson,
colonization of transgenic porcine milk proteins inhibit Hultberg et al.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
16

mice mice colonization 2006)


Helminth parasites and C-type lectin MGL recognizes (van Vliet, van
tumor antigens/ human parasite and tumor antigen Liempt et al. 2005)
dendritic cells carbohydrates
Pseudomonas MDCK mutants in saccharide (Apodaca, Bomsel
aeruginosa/ model biosynthesis are protected et al. 1995)
epithelial cells from bacteria toxicity
Pseudomonas Neuraminidase has a role in (Soong, Muir et al.
aeruginosa/ colonization mucosal infection enhancing 2006)
of the respiratory tract biofilm formation
Streptococcus Bacteria express a surface (Blau, Portnoi et al.
pneumoniae/ colonization lectin to recognize flamingo 2007)
of nasopharyngeal and cadherin receptor
lung tissues
Streptococcus Host species specificity (King, Hippe et al.
pneumoniae/human probably related with 2006)
airway epithelial cells deglycosylation of
glycoconjugates by bacterial
exoglycosidases
Streptococcus Lacto-N-neotetraose, (Tong, McIver et al.
pneumoniae/ chinchilla asialoganglioside-GM1 inhibit 1999)
tracheal epithelium adherence but neuraminidase
enhances adherence
Type I reovirus/ M cells Virus recognizes (Helander, Silvey et
interaction glycoconjugates containing al. 2003)
sialic acid in the apical
membrane of M cells
Yersinia pestis/ human Modified disaccharides affect (Thomas, Hacking
respiratory epithelial cell bacterial adhesion et al. 2007)
lines

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
17

Table 2: Host glycoproteins involved in organisms’ interactions.

Glycoprotein and Microbe Process or molecules involved Reference


Collagen
Haemophilus ducreyi Collagen-binding NcaA outer (Fulcher, Cole et
membrane protein has a role in al. 2006)
infections that lead to chancroid
(genital ulcer disease)
Staphylococcus Adhesin with a role in (Arrecubieta, Lee
epidermidis establishment of biofilms in et al. 2007)
implant devices infections
Complement C3 factor
HIV Virus opsonization enhances (Bouhlal,
infection of dendritic cells and T Chomont et al.
cells via CR3 and DC-SIGN 2007)
receptors
Klebsiella pneumoniae Adherence and internalization of (de Astorza,
bacteria into airway epithelial Cortes et al.
cells is enhanced by 2004)
opsonization
Defensin
Haemophilus influenzae Interaction with epithelial cells is (Gorter, Oostrik
and Neisseria enhanced by opsonization et al. 2003)
meningitides
Extracellular matrix
Paracoccidioides Interaction with cells through a (Barbosa, Bao et
brasiliensis surface-associated al. 2006)
glyceraldehyde-3-phosphate
dehydrogenase
Factor H
Streptococcus Invasion of mouse lungs in vivo (Quin, Onwubiko
pneumoniae is promoted by host factor H that et al. 2007)
recognizes surface bacterium
PspP receptor
Fibronectin
Bartonella henselae and Pathogen recognition of (Dabo, Confer et
Pasteurella multocida fibronectin plays a role for al. 2005; Dabo,
interaction with host Confer et al.
2006)
Clostridium difficile Adherence to epithelial cells (Hennequin,
Janoir et al.
2003)
Salmonella typhimurium Adenosine induces polarized (Walia,
fibronectin secretion by epithelial Castaneda et al.
cells and both adenosine and 2004)
fibronectin enhance bacterial
attachment and invasion

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
18

Staphylococcus aureus Truncation of fibronectin binding (Grundmeier,


protein inhibits bacterium Hussain et al.
interaction with host cells 2004)
Streptococcus A surface collagen-like protein (Caswell,
Scl1 has a role in integrins
recognition in host cells Lukomska et al.

2007)
Streptococcus agalactiae FbsA (fibronectin receptor) (Schubert,
enhances adherence to human Zakikhany et al.
epithelial cells and to human 2004;
brain microvascular endothelial Tenenbaum,
cells Bloier et al. 2005)
Streptococcus group A Fibronectin binding gene prtF2 is (Gorton, Norton
associated with enhanced et al. 2005)
bacterial internalization
Vibrio vulnificus Outer membrane OmpU has a (Goo, Lee et al.
role in microbe pathogenesis 2006)
Yersinia YadA recognition of fibronectin is (Heise and
pseudotuberculosis determinant of their ability to Dersch 2006)
infect human cells
Glycosaminoglicans
Streptococcus Attachment to nasopharyngeal (Tonnaer,
pneumoniae epithelial cells mediated by Hafmans et al.
glycosaminoglycans 2006)
Streptococcus uberis The adherence and (Almeida, Luther
internalization of the bovine et al. 2003)
mastitis-associated bacteria is
enhanced by host
glycosaminoglicans
Lactoferrin
HIV-1 Together with natural antibodies (Saidi, Eslaphazir
enhances adsorption on et al. 2006)
epithelial cells and dendritic cells
Mucin
Clostridium difficile Flagellar proteins FliC and FliD (Tasteyre, Barc
play a role in adherence and gut et al. 2001)
colonization in mice
Group A Streptococci Mucin and human pharyngeal (Ryan, Pancholi
cells are recognized through et al. 2001)
receptors containing sialic acid
Helicobacter pylori Co-localizes with MUC5AC in (Van den Brink,
human stomach Tytgat et al.
2000)
Lactobacillus johnsonii Bacterial surface-associated (Granato,
elongation factor Tu mediates Bergonzelli et al.
attachment to human intestinal 2004)

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
19

cells and mucins


Lactobacillus reuteri A prevalent gut microbe, (Miyoshi, Okada
expresses a surface protein et al. 2006)
MapA with a role in mucus and
human epithelial Caco-2 cells
recognition
Pseudomonas Mucin is recognized by adhesin- (Ramphal, Arora
aeruginosa flagellar system et al. 1996)
Staphylococcus aureus Bacterial surface proteins (Shuter, Hatcher
recognize saccharide moieties in et al. 1996)
mucin, possible role in nasal
colonization
Vibrio cholerae Adherence to mucus and villus (Yamamoto and
surface in human small intestine Yokota 1988)
Salivary CD14
Actinobacillus Invasion of human oral epithelial (Takayama,
actinomycetemcomitans cells and interleukin-8 production Satoh et al.
is enhanced 2003)
Tamm-Horsfall
Pseudomonas Urine glycoprotein enhances in (Harjai, Mittal et
aeruginosa vivo urinary infections produced al. 2005)
in mice
Thrombospondin-1
Gram-positive pathogens Adherence to host cells is (Rennemeier,
promoted through recognition of Hammerschmidt
peptidoglycan et al. 2007)
Vitronectin
Pseudomonas Enhances infection via (Leroy-Dudal,
aeruginosa integrins in airway epithelial Gagniere et al.
cells 2004)
Low-molecular-weight
human salivary mucin
Some oral bacteria Bridges interaction of bacteria (Prakobphol,
and neutrophiles Tangemann et al.
1999)
Streptococcus Surface glycoproteins GspB and (Takamatsu,
Hsa recognize mucin and sIgA Bensing et al.
2006)

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
20

References

Alaniz, M. E., R. D. Lardone, et al. (2004). "Normally Occurring Human Anti-


GM1 Immunoglobulin M Antibodies and the Immune Response to
Bacteria." Infection and Immunity 72(4): 2148-2151.
Alice, M. and S. R. Molakala (1996). "Role of Type 1 Fimbriae in the Adhesion
of Escherichia coli to Salivary Mucin and Secretory Immunoglobulin A."
Current Microbiology 33(3): 200-208.
Almeida, R. A., D. A. Luther, et al. (2003). "Binding of host glycosaminoglycans
and milk proteins: possible role in the pathogenesis of Streptococcus
uberis mastitis." Veterinary Microbiology 94(2): 131-141.
Amore, A., P. Cirina, et al. (2001). "Glycosylation of Circulating IgA in Patients
with IgA Nephropathy Modulates Proliferation and Apoptosis of
Mesangial Cells." Journal of the American Society of Nephrology 12(9):
1862-1871.
Anne Imberty, Yoann M. C. R. R. (2008). "Glycomimetics and Glycodendrimers
as High Affinity Microbial Anti-adhesins." Chemistry 14(25): 7490-7499.
Apodaca, G., M. Bomsel, et al. (1995). "Characterization of Pseudomonas
aeruginosa-induced MDCK cell injury: glycosylation-defective host cells
are resistant to bacterial killing." Infection and Immunity 63(4):
1541-1551.
Arnold, J. N., M. R. Wormald, et al. (2007). "The Impact of Glycosylation on the
Biological Function and Structure of Human Immunoglobulins." Annual
Review of Immunology 25(1): 21-50.
Arrecubieta, C., M.-H. Lee, et al. (2007). "SdrF, a Staphylococcus epidermidis
Surface Protein, Binds Type I Collagen." The Journal of Biological
Chemistry 282(26): 18767-18776.
Azab, A., J. Kleinstern, et al. (2008). "The Metastatic Stage-dependent Mucosal
Expression of Sialic Acid is a Potential Marker for Targeting Colon
Cancer with Cationic Polymers." Pharmaceutical Research 25(2):
379-386.
Balachandran, P., N. D. Pugh, et al. (2006). "Toll-like receptor 2-dependent
activation of monocytes by Spirulina polysaccharide and its immune
enhancing action in mice." International Immunopharmacology 6(12):
1808-1814.
Barbosa, M. S., S. N. Bao, et al. (2006). "Glyceraldehyde-3-Phosphate
Dehydrogenase of Paracoccidioides brasiliensis Is a Cell Surface Protein
Involved in Fungal Adhesion to Extracellular Matrix Proteins and
Interaction with Cells." Infection and Immunity 74(1): 382-389.
Barrows, B. D., J. S. Griffitts, et al. (2006). Caenorhabditis elegans
Carbohydrates in Bacterial Toxin Resistance. Methods in Enzymology,
Academic Press. Volume 417: 340-358.
Barua, D. and A. S. Paguio (1977). "ABO blood groups and cholera." Annals of
Human Biology 4(5): 489-492.
Bates, S., H. B. Hughes, et al. (2006). "Outer Chain N-Glycans Are Required for
Cell Wall Integrity and Virulence of Candida albicans." The Journal of
Biological Chemistry 281(1): 90-98.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
21

Berin, M. C., H. Li, et al. (2006). "Antibody-Mediated Antigen Sampling across


Intestinal Epithelial Barriers." Annals of the New York Academy of
Sciences 1072(1): 253-261.
Bishop, A. L., S. Wiles, et al. (2007). "Cell attachment properties and infectivity
of host-adapted and environmentally adapted Citrobacter rodentium."
Microbes and Infection 9(11): 1316-1324.
Blackwell, C. C. C. C., S. S. Dundas, et al. (2002). "Blood group and
susceptibility to disease caused by Escherichia coli O157." The Journal
of Infectious Diseases 185(3): 393-6.
Blanco, L. P. (2008). GANGLIOSIDES' ROLE IN CELLULAR INTERACTIONS
WITH MICROBES AND IMMUNE CELLS ACTIVATION. Glycolipids:
New Research. D. Sasaki. New York, Novapublishers.
Blanco, L. P. and V. J. DiRita (2006). "Bacterial-associated cholera toxin and
GM1 binding are required for transcytosis of classical biotype Vibrio
cholerae through an in vitro M cell model system." Cellular Microbiology
8(6): 982-998.
Blanco, L. P. and V. J. DiRita (2006a). "Antibodies Enhance Interaction of Vibrio
cholerae with Intestinal M-Like Cells." Infection and Immunity 74(12):
6957-6964.
Blau, K., M. Portnoi, et al. (2007). "Flamingo Cadherin: A Putative Host
Receptor for Streptococcus pneumoniae." The Journal of Infectious
Diseases 195(12): 1828-1837.
Bouhlal, H., N. Chomont, et al. (2007). "Opsonization of HIV with Complement
Enhances Infection of Dendritic Cells and Viral Transfer to CD4 T Cells in
a CR3 and DC-SIGN-Dependent Manner." Journal of Immunology
178(2): 1086-1095.
Brennan, M. J., J. O. Cisar, et al. (1984). "Lectin-dependent attachment of
Actinomyces naeslundii to receptors on epithelial cells." Infection and
Immunity 46(2): 459-464.
Brennan, M. J., R. A. Joralmon, et al. (1987). "Binding of Actinomyces
naeslundii to glycosphingolipids." Infection and Immunity 55(2): 487-489.
Brewer, C. F., M. C. Miceli, et al. (2002). "Clusters, bundles, arrays and lattices:
novel mechanisms for lectin-saccharide-mediated cellular interactions."
Current Opinion in Structural Biology 12(5): 616-623.
Brockhausen, I. (2003). "Glycodinamics of mucin biosynthesis in
gastrointestinal tumor cells." Advanced Experimental Medical Biology
535: 163-188.
Bucior, I. and M. Burger (2004). "Carbohydrate-carbohydrate interaction as a
major force initiating cell-cell recognition." Glycoconjugate Journal 21(3):
111-123.
Bucior, I. and M. M. Burger (2004). "Carbohydrate-carbohydrate interactions in
cell recognition." Current Opinion in Structural Biology 14(5): 631-637.
Bucior, I., S. Scheuring, et al. (2004). "Carbohydrate-carbohydrate interaction
provides adhesion force and specificity for cellular recognition." Journal
of Cell Biology 165(4): 529-537.
Buzás, E. I. E. I., B. B. György, et al. (2006). "Carbohydrate recognition systems
in autoimmunity." Autoimmunity 39(8): 691-704.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
22

Cao, Z., D. M. Jefferson, et al. (1998). "Role of Carbohydrate-mediated


Adherence in Cytopathogenic Mechanisms of Acanthamoeba." The
Journal of Biological Chemistry 273(25): 15838-15845.
Casadevall, A. and L.-a. Pirofski (2003). "The damage-response framework of
microbial pathogenesis." Nat Rev Micro 1(1): 17-24.
Caswell, C. C., E. Lukomska, et al. (2007). "Scl1-dependent internalization of
group A Streptococcus via direct interactions with the
α2β1 integrin enhances pathogen survival and re-
emergence." Molecular Immunology 64(5): 1319-1331.
Cipollo, J. F., A. M. Awad, et al. (2004). "srf-3, a Mutant of Caenorhabditis
elegans, Resistant to Bacterial Infection and to Biofilm Binding, Is
Deficient in Glycoconjugates." The Journal of Biological Chemistry
279(51): 52893-52903.
Cisneros, R. L., F. C. Gibson, 3rd, et al. (1996). "Passive transfer of poly-(1-6)-
beta-glucotriosyl-(1-3)-beta- glucopyranose glucan protection against
lethal infection in an animal model of intra-abdominal sepsis." Infection
and Immunity 64(6): 2201-2205.
Cormack, B. P., N. Ghori, et al. (1999). "An Adhesin of the Yeast Pathogen
Candida glabrata Mediating Adherence to Human Epithelial Cells."
Science 285(5427): 578-582.
Corthesy, B., M. Kaufmann, et al. (1996). "A Pathogen-specific Epitope Inserted
into Recombinant Secretory Immunoglobulin A Is Immunogenic by the
Oral Route." The Journal of Biological Chemistry 271(52): 33670-33677.
Dabo, S. M., A. W. Confer, et al. (2006). "Bartonella henselae Pap31, an
Extracellular Matrix Adhesin, Binds the Fibronectin Repeat III13 Module."
Infection and Immunity 74(5): 2513-2521.
Dabo, S. M., A. W. Confer, et al. (2005). "Adherence of Pasteurella multocida to
fibronectin." Veterinary Microbiology 110(3-4): 265-275.
Dall'Olio, F. and M. Chiricolo (2001). "Sialyltransferases in cancer."
Glycoconjugate Journal 18(11): 841-850.
de Astorza, B., G. Cortes, et al. (2004). "C3 Promotes Clearance of Klebsiella
pneumoniae by A549 Epithelial Cells." Infection and Immunity 72(3):
1767-1774.
De Praeter, C., G. J. Gerwig, et al. (2000). "A novel disorder caused by
defective biosynthesis of N-linked oligosaccharides due to glucosidase I
deficiency." American Journal Human Genetics 66: 1744-1756.
Eklund, E. A. and H. H. Freeze (2005). "Essentials of Glycosylation." Seminars
in Pediatric Neurology 12(3): 134-143.
Evans, S. V. and R. MacKenzie (1999). "Characterization of protein-glycolipid
recognition at the membrane bilayer." Journal of Molecular Recognition
12: 155-168.
Faruque, S. M., K. Biswas, et al. (2006). "Transmissibility of cholera: In vivo-
formed biofilms and their relationship to infectivity and persistence in the
environment." Proceedings of the National Academy of Sciences
103(16): 6350-6355.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
23

Favre, L., F. Spertini, et al. (2005). "Secretory IgA Possesses Intrinsic


Modulatory Properties Stimulating Mucosal and Systemic Immune
Responses." Journal of Immunology 175(5): 2793-2800.
Flick, K., C. Scholander, et al. (2001). "Role of Nonimmune IgG Bound to
PfEMP1 in Placental Malaria." Science 293(5537): 2098-2100.
Freeze, H. H. (2001). "Update and perspectives on congenital disorders of
glycosylation." Glycobiology 11(12): 129R-143.
Freeze, H. H. (2002). "Human disorders in N-glycosylation and animal models."
Biochimica et Biophysica Acta (BBA) - General Subjects 1573(3):
388-393.
Freeze, H. H. (2006). "Genetic defects in the human glycome." Nat Rev Genet
7(7): 537-551.
Fulcher, R. A., L. E. Cole, et al. (2006). "Expression of Haemophilus ducreyi
Collagen Binding Outer Membrane Protein NcaA Is Required for
Virulence in Swine and Human Challenge Models of Chancroid."
Infection and Immunity 74(5): 2651-2658.
Gao, Y. H., L. X. Xu, et al. (2007). "Differential binding characteristics of native
monomeric and polymeric immunoglobulin A1 (IgA1) on human
mesangial cells and the influence of in vitro deglycosylation of IgA1
molecules." Clinical Experimental Immunology 148(3): 507-514.
Garate, M., J. Marchant, et al. (2006). "In Vitro Pathogenicity of Acanthamoeba
Is Associated with the Expression of the Mannose-Binding Protein."
Investigative Ophthalmology and Visual Science 47(3): 1056-1062.
Garcia Guijo, C., A. Garcia-Merino, et al. (1995). "Presence and isotype of anti-
ganglioside antibodies in healthy persons, motor neuron disease,
peripheral neuropathy, and other diseases of the nervous system."
Journal of Neuroimmunology 56(1): 27-33.
Golks, A., T.-T. T. Tran, et al. (2007). "Requirement for O-linked N-
acetylglucosaminyltransferase in lymphocytes activation." The EMBO
Journal 26(20): 4368-79.
Goo, S. Y., H.-J. Lee, et al. (2006). "Identification of OmpU of Vibrio vulnificus
as a Fibronectin-Binding Protein and Its Role in Bacterial Pathogenesis."
Infection and Immunity 74(10): 5586-5594.
Gorter, A. D., J. Oostrik, et al. (2003). "Involvement of lipooligosaccharides of
Haemophilus influenzae and Neisseria meningitidis in defensin-
enhanced bacterial adherence to epithelial cells." Microbial Pathogenesis
34(3): 121-130.
Gorton, D., R. Norton, et al. (2005). "Presence of fibronectin-binding protein
gene prtF2 in invasive group A streptococci in tropical Australia is
associated with increased internalisation efficiency." Microbes and
Infection 7(3): 421-426.
Gradimir, N. M. (1999). "Molecular self-recognition and adhesion via
proteoglycan to proteoglycan interactions as a pathway to multicellularity:
Atomic force microscopy and color coded bead measurements in
sponges." Microscopy Research and Technique 44(4): 304-309.
Granato, D. D., G. E. G. E. Bergonzelli, et al. (2004). "Cell surface-associated
elongation factor Tu mediates the attachment of Lactobacillus johnsonii

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
24

NCC533 (La1) to human intestinal cells and mucins." Infection and


Immunity 72(4): 2160-9.
Griffitts, J. S., S. M. Haslam, et al. (2005). "Glycolipids as Receptors for Bacillus
thuringiensis Crystal Toxin." Science 307(5711): 922-925.
Griffitts, J. S., D. L. Huffman, et al. (2003). "Resistance to a Bacterial Toxin Is
Mediated by Removal of a Conserved Glycosylation Pathway Required
for Toxin-Host Interactions." The Journal of Biological Chemistry 278(46):
45594-45602.
Grundmeier, M., M. Hussain, et al. (2004). "Truncation of Fibronectin-Binding
Proteins in Staphylococcus aureus Strain Newman Leads to Deficient
Adherence and Host Cell Invasion Due to Loss of the Cell Wall Anchor
Function." Infection and Immunity 72(12): 7155-7163.
Gustafsson, A., A. Hultberg, et al. (2006). "Carbohydrate-dependent inhibition of
Helicobacter pylori colonization using porcine milk." Glycobiology 16(1):
1-10.
Haghighi, H. R., J. Gong, et al. (2006). "Probiotics Stimulate Production of
Natural Antibodies in Chickens." Clinical Vaccine Immunology 13(9):
975-980.
Hakomori, S.-i. (2002). "Inaugural Article: The glycosynapse." Proceedings of
the National Academy of Sciences 99(1): 225-232.
Hakomori, S. (2004). "Carbohydrate-to-carbohydrate interaction, through
glycosynapse, as a basis of cell recognition and membrane
organization." Glycoconjugate Journal 21(3): 125-137.
Hakomori, S. (2004b). "Glycosynapses: microdomains controlling carbohydrate-
dependent cell adhesion and signalling." Annals Brazilian Acad Sci 76:
553-572.
Hansen, G. H., E. D. K. Pedersen, et al. (2005). "Anti-glycosyl antibodies in lipid
rafts of the enterocyte brush border: a possible host defense against
pathogens." American Journal Gastrointestinal and Liver Physiology
289(6): G1100-1107.
Harjai, K., R. Mittal, et al. (2005). "Contribution of Tamm-Horsfall protein to
virulence of Pseudomonas aeruginosa in urinary tract infection."
Microbes and Infection 7(1): 132-137.
Haseley, S. R., H. J. Vermeer, et al. (2001). "Carbohydrate self-recognition
mediates marine sponge cellular adhesion." Proceedings of the National
Academy of Sciences 98(16): 9419-9424.
Heise, T. and P. Dersch (2006). "Identification of a domain in Yersinia virulence
factor YadA that is crucial for extracellular matrix-specific cell adhesion
and uptake." Proceedings of the National Academy of Sciences 103(9):
3375-3380.
Helander, A., K. J. Silvey, et al. (2003). "The Viral {sigma}1 Protein and
Glycoconjugates Containing {alpha}2-3-Linked Sialic Acid Are Involved in
Type 1 Reovirus Adherence to M Cell Apical Surfaces." Journal of
Virology 77(14): 7964-7977.
Hennequin, C., C. Janoir, et al. (2003). "Identification and characterization of a
fibronectin-binding protein from Clostridium difficile." Microbiology
149(10): 2779-2787.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
25

Henry, S. M. (2001). "Molecular diversity in the biosynthesis of GI tract


glycoconjugates. A blood group related chart of microorganism
receptors." Transfusion Clinique et Biologique 8(3): 226-230.
Höflich, J., P. Berninsone, et al. (2004). "Loss of srf-3-encoded Nucleotide
Sugar Transporter Activity in Caenorhabditis elegans Alters Surface
Antigenicity and Prevents Bacterial Adherence." Journal of Biological
Chemistry 279(29): 30440-30448.
Hollingsworth, M. A. and B. J. Swanson (2004). "Mucins in cancer: protection
and control of the cell surface." Nat Rev Cancer 4(1): 45-60.
Hsu, K.-L., K. T. Pilobello, et al. (2006). "Analyzing the dynamic bacterial
glycome with a lectin microarray approach." Nat Chem Biol 2(3):
153-157.
Hutchings, A. B., A. Helander, et al. (2004). "Secretory Immunoglobulin A
Antibodies against the {sigma}1 Outer Capsid Protein of Reovirus Type 1
Lang Prevent Infection of Mouse Peyer's Patches." Journal of Virology
78(2): 947-957.
Hutson, A. M., R. L. Atmar, et al. (2002). "Norwalk virus infection and disease is
associated with ABO histo-blood group type." Journal of Infectious
Diseases 185(9): 1335-1337.
Jiang, H. Q., M. C. Thurnheer, et al. (2004). "Interactions of commensal gut
microbes with subsets of B- and T-cells in the murine host." Vaccine
22(7): 805-811.
Jost, B. H. B. H., J. G. J. G. Songer, et al. (2002). "Identification of a second
Arcanobacterium pyogenes neuraminidase and involvement of
neuraminidase activity in host cell adhesion." Infection and Immunity
70(3): 1106-12.
Karjalainen, T., M. C. Barc, et al. (1994). "Cloning of a genetic determinant from
Clostridium difficile involved in adherence to tissue culture cells and
mucus." Infection and Immunity 62(10): 4347-4355.
Kerr, M. A. (2000). "Function of immunoglobulin A in immunity." Gut 47(6):
751-752.
Kijima-Suda, I., Y. Miyamoto, et al. (1986). "Inhibition of Experimental
Pulmonary Metastasis of Mouse Colon Adenocarcinoma 26 Sublines by
a Sialic Acid:Nucleoside Conjugate Having Sialyltransferase Inhibiting
Activity." Cancer Research 46(2): 858-862.
Kijima-Suda, I., T. Miyazawa, et al. (1988). "Possible Mechanism of Inhibition of
Experimental Pulmonary Metastasis of Mouse Colon Adenocarcinoma 26
Sublines by a Sialic Acid:Nucleoside Conjugate." Cancer Research
48(13): 3728-3732.
King, S. J., K. R. Hippe, et al. (2006). "Deglycosylation of human
glycoconjugates by the sequential activities of exoglycosidases
expressed by Streptococcus pneumoniae." Molecular Microbiology 59(3):
961-974.
Krivan, H. C., D. D. Roberts, et al. (1988). "Many Pulmonary Pathogenic
Bacteria Bind Specifically to the Carbohydrate Sequence GalNAc{beta}
1-4Gal Found in Some Glycolipids." Proceedings of the National
Academy of Sciences 85(16): 6157-6161.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
26

Kroese, F. G. M., R. de Waard, et al. (1996). "B-1 cells and their reactivity with
the murine intestinal microflora." Seminars in Immunology 8(1): 11-18.
Landers, J. J., Z. Cao, et al. (2002). "Prevention of influenza pneumonitis by
sialic Acid-conjugated dendritic polymers." The Journal of Infectious
Diseases 186(9): 1222-30.
Lauc, G. and M. Heffer-Lauc (2006). "Shedding and uptake of gangliosides and
glycosylphosphatidylinositol-anchored proteins." Biochimica et
Biophysica Acta (BBA) - General Subjects 1760(4): 584-602.
Leroy-Dudal, J., H. Gagniere, et al. (2004). "Role of [alpha]v[beta]5 integrins
and vitronectin in Pseudomonas aeruginosa PAK interaction with A549
respiratory cells." Microbes and Infection 6(10): 875-881.
Li, H., A. Nowak-Wegrzyn, et al. (2006). "Transcytosis of IgE-Antigen
Complexes by CD23a in Human Intestinal Epithelial Cells and Its Role in
Food Allergy." Gastroenterology 131(1): 47-58.
Liu, D., C. C. Cramer, et al. (2005). "N-Linked Glycosylation at Asn3 and the
Positively Charged Residues within the Amino-Terminal Domain of the
C1 Inhibitor Are Required for Interaction of the C1 Inhibitor with
Salmonella enterica Serovar Typhimurium Lipopolysaccharide and Lipid
A." Infection and Immunity 73(8): 4478-4487.
Macpherson, A. J., L. Hunziker, et al. (2001). "IgA responses in the intestinal
mucosa against pathogenic and non-pathogenic microorganisms."
Microbes and Infection 3(12): 1021-1035.
Macpherson, A. J. and T. Uhr (2004). "Induction of Protective IgA by Intestinal
Dendritic Cells Carrying Commensal Bacteria." Science 303(5664):
1662-1665.
Magnusson, K. K. E. and I. I. Stjernström (1982). "Mucosal barrier mechanisms.
Interplay between secretory IgA (SIgA), IgG and mucins on the surface
properties and association of salmonellae with intestine and
granulocytes." Immunology 45(2): 239-48.
Manes, S., G. del Real, et al. (2003). "Pathogens: raft hijackers." Nat Rev
Immunol 3(7): 557-568.
Martin, P. T. and H. H. Freeze (2003). "Glycobiology of neuromuscular
disorders." Glycobiology 13(8): 67R-75.
McCool, T. L. and J. N. Weiser (2004). "Limited Role of Antibody in Clearance
of Streptococcus pneumoniae in a Murine Model of Colonization."
Infection and Immunity 72(10): 5807-5813.
Mentis, A., C. C. Blackwell, et al. (1991). "ABO blood group, secretor status and
detection of Helicobacter pylori among patients with gastric or duodenal
ulcers." Epidemiology and Infectology 106: 221-229.
Merrell, D. S., S. M. Butler, et al. (2002). "Host-induced epidemic spread of the
cholera bacterium." Nature 417(6889): 642-645.
Mestecky, J., M. W. Russell, et al. (1999). "Intestinal IgA: novel views on its
function in the defence of the largest mucosal surface." Gut 44(1): 2-5.
Miller, B. S. and H. H. Freeze (2003). "New Disorders in Carbohydrate
Metabolism: Congenital Disorders of Glycosylation and Their Impact on
the Endocrine System." Reviews in Endocrine & Metabolic Disorders
4(1): 103-113.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
27

Misevic, G. N., Y. Guerardel, et al. (2004). "Molecular Recognition between


Glyconectins as an Adhesion Self-assembly Pathway to Multicellularity."
The Journal of Biological Chemistry 279(15): 15579-15590.
Miyoshi, Y., S. Okada, et al. (2006). "A mucus adhesion promoting protein,
MapA, mediates the adhesion of Lactobacillus reuteri to Caco-2 human
intestinal epithelial cells." Bioscience, Biotechnology, and Biochemistry
70(7): 1622-8.
Monks, C. R. F., B. A. Freiberg, et al. (1998). "Three-dimensional segregation of
supramolecular activation clusters in T cells." Nature 395(6697): 82-86.
Munro, C. A., S. Bates, et al. (2005). "Mnt1p and Mnt2p of Candida albicans
Are Partially Redundant {alpha}-1,2-Mannosyltransferases That
Participate in O-Linked Mannosylation and Are Required for Adhesion
and Virulence." The Journal of Biological Chemistry 280(2): 1051-1060.
Nashar, T. O., Z. E. Betteridge, et al. (2002). "Antigen binding to GM1
ganglioside results in delayed presentation: minimal effects of GM1 on
presentation of antigens internalized via other pathways." Immunology
106(1): 60-70.
Netea, M. G., N. A. R. Gow, et al. (2006). "Immune sensing of Candida albicans
requires cooperative recognition of mannans and glucans by lectin and
Toll-like receptors." Journal of Clinical Investigation 116(6): 1642-1650.
Ohashi, Y., M. Hiraguchi, et al. (2006). "The composition of intestinal bacteria
affects the level of luminal IgA." Bioscience, Biotechnology, and
Biochemistry 70(12): 3031-5.
Ono, M. and S. Hakomori (2003). "Glycosylation defining cancer cell motility
and invasiveness." Glycoconjugate Journal 20(1): 71-78.
Palestrant, D., Z. E. Holzknecht, et al. (2004). "Microbial Biofilms in the Gut:
Visualization by Electron Microscopy and by Acridine Orange Staining."
Ultrastructural Pathology 28(1): 23 - 27.
Paulsen, H. and I. Brockhausen (2001). "From imino sugars to cancer
glycoproteins." Glycoconjugate Journal 18(11): 867-870.
Perrier, C., N. Sprenger, et al. (2006). "Glycans on Secretory Component
Participate in Innate Protection against Mucosal Pathogens." The Journal
of Biological Chemistry 281(20): 14280-14287.
Popescu, O., I. Checiu, et al. (2003). "Quantitative and qualitative approach of
glycan-glycan interactions in marine sponges." Biochimie 85(1-2):
181-188.
Prakobphol, A., K. Tangemann, et al. (1999). "Separate Oligosaccharide
Determinants Mediate Interactions of the Low-Molecular-Weight Salivary
Mucin with Neutrophils and Bacteria." Biochemistry 38(21): 6817-6825.
Quin, L. R., C. Onwubiko, et al. (2007). "Factor H Binding to PspC of
Streptococcus pneumoniae Increases Adherence to Human Cell Lines In
Vitro and Enhances Invasion of Mouse Lungs In Vivo." Infection and
Immunity 75(8): 4082-4087.
Rabin, S. J., A. Bachis, et al. (2002). "Gangliosides Activate Trk Receptors by
Inducing the Release of Neurotrophins." Journal of Biological Chemistry
277(51): 49466-49472.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
28

Ramphal, R., S. K. Arora, et al. (1996). "Recognition of mucin by the adhesin-


flagellar system of Pseudomonas aeruginosa." American Journal of
Respiratory and Critical Care Medicine 154(4 Pt 2): 4.
Rennemeier, C., S. Hammerschmidt, et al. (2007). "Thrombospondin-1
promotes cellular adherence of Gram-positive pathogens via recognition
of peptidoglycan." the FASEB Journal 21(12): 3118-3132.
Restrepo, C. I., Q. Dong, et al. (1999). "Surfactant Protein D Stimulates
Phagocytosis of Pseudomonas aeruginosa by Alveolar Macrophages."
American Journal of Respiratory Cell and Molecular Biology 21(5):
576-585.
Robbe, C., C. Capon, et al. (2004). "Structural diversity and specific distribution
of O-glycans in normal human mucins along the intestinal tract."
Biochemistry Journal 384(2): 307-316.
Royle, L., A. Roos, et al. (2003). "Secretory IgA N- and O-Glycans Provide a
Link between the Innate and Adaptive Immune Systems." The Journal of
Biological Chemistry 278(22): 20140-20153.
Rudd, P. M., M. R. Wormald, et al. (2004). "Sugar-mediated ligand-receptor
interactions in the immune system." Trends in Biotechnology 22(10):
524-530.
Ryan, P. A., V. Pancholi, et al. (2001). "Group A Streptococci Bind to Mucin and
Human Pharyngeal Cells through Sialic Acid-Containing Receptors."
Infection and Immunity 69(12): 7402-7412.
Sadighi Akha, A. A., S. B. Berger, et al. (2006). "Enhancement of CD8 T-cell
function through modifying surface glycoproteins in young and old mice."
immunology 119(2): 187-194.
Sadighi Akha, A. A. and R. A. Miller (2005). "Signal transduction in the aging
immune system." Current Opinion in Immunology 17(5): 486-491.
Saidi, H., J. Eslaphazir, et al. (2006). "Differential Modulation of Human
Lactoferrin Activity against Both R5 and X4-HIV-1 Adsorption on
Epithelial Cells and Dendritic Cells by Natural Antibodies." J Immunol
177(8): 5540-5549.
Saito, T. and T. Yokosuka (2006). "Immunological synapse and microclusters:
the site for recognition and activation of T cells." Current Opinion in
Immunology 18(3): 305-313.
Sajjan, U., J. Moreira, et al. (2004). "A novel model to study bacterial adherence
to the transplanted airway: inhibition of Burkholderia cepacia adherence
to human airway by dextran and xylitol." The Journal of Heart and Lung
Transplantation 23(12): 1382-91.
Sandini, S., R. La Valle, et al. (2007). "The 65 kDa mannoprotein gene of
Candida albicans encodes a putative β-glucanase adhesin
required for hyphal morphogenesis and experimental pathogenicity."
Cellular Microbiology 9(5): 1223-1238.
Schengrund, C.-L. (2003). ""Multivalent" saccharides: development of new
approaches for inhibiting the effects of glycosphingolipid-binding
pathogens." Biochemical Pharmacology 65(5): 699-707.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
29

Schubert, A., K. Zakikhany, et al. (2004). "The Fibrinogen Receptor FbsA


Promotes Adherence of Streptococcus agalactiae to Human Epithelial
Cells." Infection and Immunity 72(11): 6197-6205.
Schwarzmann, G. (2001). "Uptake and metabolism of exogenous
glycosphingolipids by cultured cells." Seminars in Cell & Developmental
Biology 12(2): 163-171.
Sein, J., V. Cachicas, et al. (1993). "Mucins allow survival of Salmonella typhi
within mouse peritoneal macrophages." Biological Research 26:
371-380.
Shen, W., R. Falahati, et al. (2005). "Modulation of CD4 Th Cell Differentiation
by Ganglioside GD1a In Vitro." Journal of Immunology 175(8):
4927-4934.
Shoaf, K., G. L. Mulvey, et al. (2006). "Prebiotic Galactooligosaccharides
Reduce Adherence of Enteropathogenic Escherichia coli to Tissue
Culture Cells." Infection and Immunity 74(12): 6920-6928.
Shuter, J., V. B. Hatcher, et al. (1996). "Staphylococcus aureus binding to
human nasal mucin." Infection and Immunity 64(1): 310-8.
Singh, A. P., N. Moniaux, et al. (2004). "Inhibition of MUC4 expression
suppresses pancreatic tumor cell growth and metastasis." Cancer
Research 64(2): 622-30.
Sonnenburg, J. L., L. T. Angenent, et al. (2004). "Getting a grip on things: how
do communities of bacterial symbionts become established in our
intestine?" Nature Immunology 5(6): 569-73.
Sonnino, S., L. Mauri, et al. (2007). "Gangliosides as components of lipid
membrane domains." Glycobiology 17(1): 1R-13.
Soong, G., A. Muir, et al. (2006). "Bacterial neuraminidase facilitates mucosal
infection by participating in biofilm production." J. Clin. Invest. 116(8):
2297-2305.
Spiegel, S., J. Schlessinger, et al. (1984). "Incorporation of fluorescent
gangliosides into human fibroblasts: mobility, fate, and interaction with
fibronectin." Journal of Cell Biology 99(2): 699-704.
Springer, G. F. (1970). "Importance of blood-group substances in interactions
between man and microbes." Annals of the New York Academy of
Sciences 169(1): 134-152.
Stoel, M., W. N. H. Evenhuis, et al. (2008). "Rat salivary gland reveals a more
restricted IgA repertoire than ileum." Molecular Immunology 45(3):
719-727.
Stoel, M., H.-Q. Jiang, et al. (2005). "Restricted IgA Repertoire in Both B-1 and
B-2 Cell-Derived Gut Plasmablasts." Journal of Immunology 174(2):
1046-1054.
Stuart, E. S., W. C. Webley, et al. (2003). "Lipid rafts, caveolae, caveolin-1, and
entry by Chlamydiae into host cells." Experimental Cell Research 287(1):
67-78.
Sugii, S. (1990). "Binding specificity of the combining site of cholera toxin for
human erythrocytes." Canadian Journal of Microbiology 36(6): 452-454.
Suzuki, M., M. Suzuki, et al. (2005). "Polysialic acid facilitates tumor invasion by
glioma cells." Glycobiology 15(9): 887-894.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
30

Takada, A., K. Fujioka, et al. (2004). "Human Macrophage C-Type Lectin


Specific for Galactose and N-Acetylgalactosamine Promotes Filovirus
Entry." Journal of Virology 78(6): 2943-2947.
Takamatsu, D., B. A. Bensing, et al. (2006). "Binding of the Streptococcal
Surface Glycoproteins GspB and Hsa to Human Salivary Proteins."
Infection and Immunity 74(3): 1933-1940.
Takayama, A., A. Satoh, et al. (2003). "Augmentation of Actinobacillus
actinomycetemcomitans Invasion of Human Oral Epithelial Cells and Up-
Regulation of Interleukin-8 Production by Saliva CD14." Infection and
Immunity 71(10): 5598-5604.
Tasteyre, A., M.-C. Barc, et al. (2001). "Role of FliC and FliD Flagellar Proteins
of Clostridium difficile in Adherence and Gut Colonization." Infection and
Immunity 69(12): 7937-7940.
Tenenbaum, T., C. Bloier, et al. (2005). "Adherence to and Invasion of Human
Brain Microvascular Endothelial Cells Are Promoted by Fibrinogen-
Binding Protein FbsA of Streptococcus agalactiae." Infection and
Immunity 73(7): 4404-4409.
Thomas, R. and T. Brooks (2004). "Common oligosaccharide moieties inhibit
the adherence of typical and atypical respiratory pathogens." Journal of
Medical Microbiology 53(9): 833-840.
Thomas, R. J. and T. J. Brooks (2004). "Oligosaccharide receptor mimics inhibit
Legionella pneumophila attachment to human respiratory epithelial cells."
Microbial Pathogenesis 36(2): 83-92.
Thomas, R. J., A. Hacking, et al. (2007). "Structural modification of a base
disaccharide alters antiadhesion properties towards Yersinia pestis."
FEMS Immunology and Medical Microbiology 49(3): 410-414.
Tong, H. H., M. A. McIver, et al. (1999). "Effect of lacto-N-neotetraose,
asialoganglioside-GM1 and neuraminidase on adherence of otitis media-
associated serotypes ofStreptococcus pneumoniaeto chinchilla tracheal
epithelium." Microbial Pathogenesis 26(2): 111-119.
Tonnaer, E. L. G. M., T. G. Hafmans, et al. (2006). "Involvement of
glycosaminoglycans in the attachment of pneumococci to
nasopharyngeal epithelial cells." Microbes and Infection 8(2): 316-322.
Tsuboi, S. and M. Fukuda (1998). "Overexpression of Branched O-Linked
Oligosaccharides on T Cell Surface Glycoproteins Impairs Humoral
Immune Responses in Transgenic Mice." Journal of Biological Chemistry
273(46): 30680-30687.
Tsuboi, S. and M. Fukuda (2001). "Roles of <I>O</I>-linked oligosaccharides in
immune responses." Bioessays 23(1): 46-53.
Tsutsumida, H., B. J. Swanson, et al. (2006). "RNA Interference Suppression of
MUC1 Reduces the Growth Rate and Metastatic Phenotype of Human
Pancreatic Cancer Cells." Clinical Cancer Research 12(10): 2976-2987.
Tzianabos, A. O. (2000). "Polysaccharide Immunomodulators as Therapeutic
Agents: Structural Aspects and Biologic Function." Clinical Microbiology
Reviews 13(4): 523-533.
van de Wetering, J. K., F. E. J. Coenjaerts, et al. (2004). "Aggregation of
Cryptococcus neoformans by Surfactant Protein D Is Inhibited by Its

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
31

Capsular Component Glucuronoxylomannan." Infection and Immunity


72(1): 145-153.
Van den Brink, G. G. R., K. K. M. Tytgat, et al. (2000). "H pylori colocalises with
MUC5AC in the human stomach." Gut 46(5): 601-7.
van Vliet, S. J., E. van Liempt, et al. (2005). "Carbohydrate profiling reveals a
distinctive role for the C-type lectin MGL in the recognition of helminth
parasites and tumor antigens by dendritic cells." International
Immunology 17(5): 661-669.
Vigerust, D. J. and V. L. Shepherd (2007). "Virus glycosylation: role in virulence
and immune interactions." Trends in Microbiology 15(5): 211-218.
Vollmers, H. P. and S. Brandlein (2006). "Natural IgM antibodies: The orphaned
molecules in immune surveillance." Advanced Drug Delivery Reviews
58(5-6): 755-765.
Vollmers, H. P. and S. Brandlein (2007). "Natural antibodies and cancer."
Journal of Autoimmunity 29(4): 295-302.
Wagner, H. H. E., P. P. Thomas, et al. (1990). "Inhibition of sialic acid
incorporation prevents hepatic metastases." Archives of Surgery 125(3):
351-4.
Walia, B., F. E. Castaneda, et al. (2004). "Polarized fibronectin secretion
induced by adenosine regulates bacterial-epithelial interaction in human
intestinal epithelial cells." Biochemical Journal 382(2): 589-596.
Walsh, M. D., J. P. Young, et al. (2007). "The MUC13 cell surface mucin is
highly expressed by human colorectal carcinomas." Human Pathology
38(6): 883-892.
Watson, P. R., S. M. Paulin, et al. (2000). "Interaction of Salmonella serotypes
with porcine macrophages in vitro does not correlate with virulence."
Microbiology 146(7): 1639-1649.
Wellens, A., C. Garofalo, et al. (2008). "Intervening with Urinary Tract Infections
Using Anti-Adhesives Based on the Crystal Structure of the
FimH–Oligomannose-3 Complex." PLoS ONE 3(4): e2040.
Wiles, S. S., G. G. Dougan, et al. (2005). "Emergence of a 'hyperinfectious'
bacterial state after passage of Citrobacter rodentium through the host
gastrointestinal tract." Cellular Microbiology 7(8): 1163-72.
Wold, A. E., J. Mestecky, et al. (1990). "Secretory immunoglobulin A carries
oligosaccharide receptors for Escherichia coli type 1 fimbrial lectin."
Infection and Immunity 58(9): 3073-3077.
Yamamoto, T. T. and T. T. Yokota (1988). "Electron microscopic study of Vibrio
cholerae O1 adherence to the mucus coat and villus surface in the
human small intestine." Infection and Immunity 56(10): 2753-9.
Yang, Z., Z. Cao, et al. (1997). "Pathogenesis of Acanthamoeba keratitis:
carbohydrate-mediated host- parasite interactions." Infection and
Immunity 65(2): 439-445.
Zaas, D. W., M. Duncan, et al. (2005). "The role of lipid rafts in the
pathogenesis of bacterial infections." Biochimica et Biophysica Acta
(BBA) - Molecular Cell Research 1746(3): 305-313.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology
32

Zaleski, K. J., T. Kolodka, et al. (2006). "Hyaluronic Acid Binding Peptides


Prevent Experimental Staphylococcal Wound Infection." Antimicrobial
Agents and Chemotherapy 50(11): 3856-3860.
Zhang, P., S. Snyder, et al. (2006). "Role of N-Acetylglucosamine within Core
Lipopolysaccharide of Several Species of Gram-Negative Bacteria in
Targeting the DC-SIGN (CD209)." Journal of Immunology 177(6):
4002-4011.

Luz P. Blanco Ph.D. (March, 2009)


Light White Innovation Technology

You might also like