Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Experimental Parasitology 126 (2010) 283–291

Contents lists available at ScienceDirect

Experimental Parasitology
journal homepage: www.elsevier.com/locate/yexpr

Molecular mechanisms of host cell invasion by Trypanosoma cruzi


Conrad L. Epting a,*, Bria M. Coates a, David M. Engman b
a
Department of Pediatrics, Northwestern University, Chicago, IL, United States
b
Departments of Pathology and Microbiology–Immunology, Northwestern University, Chicago, IL, United States

a r t i c l e i n f o a b s t r a c t

Article history: The protozoan parasite Trypanosoma cruzi, the etiologic agent of Chagas disease, is an obligate intracel-
Received 13 November 2009 lular protozoan pathogen. Overlapping mechanisms ensure successful infection, yet the relationship
Received in revised form 28 May 2010 between these cellular events and clinical disease remains obscure. This review explores the process of
Accepted 14 June 2010
cell invasion from the perspective of cell surface interactions, intracellular signaling, modulation of the
Available online 18 June 2010
host cytoskeleton and endosomal compartment, and the intracellular innate immune response to
infection.
Keywords:
Ó 2010 Elsevier Inc. All rights reserved.
Trypanosoma cruzi
Invasion
Lysosome
Microtubules
Signaling

1. Introduction explore the process of cell invasion with a focus on known cell sur-
face interactions, review the evidence surrounding tissue tropism,
The protozoan parasite Trypanosoma cruzi is the etiologic agent explore the intracellular response to infection, and highlight sev-
of Chagas disease, a disorder of poverty endemic to Central and eral experimental unknowns and challenges.
South America. Ten to sixteen million people are chronically in- T. cruzi has developed complex and redundant mechanisms to
fected with T. cruzi, with widely variable clinical sequelae, ranging ensure successful cell invasion. We will examine the common fea-
from no disease (the majority) to an inflammatory cardiomyopathy tures that underlie critical events involved in host cell infection by
and dilatation of the enteric viscera from denervation injury (Coura the parasite, with a focus on the trypomastigote. Considerable het-
and Dias, 2009; Coura, 2007; Moncayo and Silveira, 2009). Chagas erogeneity exists at each step of this process, and the specifics may
disease is emerging in North America in animals and humans, vary with each unique combination of parasite strain, stage, and
likely from the economic migration of infected individuals and host cell. Therefore, the reader is cautioned not to apply a reduc-
the extending range of the insect vector (Milei et al., 2009; Schmu- tionistic viewpoint broadly. Further, the outcome of T. cruzi infec-
nis, 2007). The early diagnosis and treatment of Chagas disease re- tion is highly heterogeneous across cell types. Aspects of cell
mains a challenge for resource-poor nations, with the acute phase invasion which vary across cell types include surface–surface
often passing undetected, and therapy during the chronic phase interactions, enzymatic events, calcium-mediated signaling, traf-
being largely supportive rather than curative (Rassi et al., 2009; ficking of donor and host membranes, cytoskeletal contributions
Perez-Molina et al., 2009). to parasite uptake and, finally, cytoplasmic entry via escape from
Typically, T. cruzi infection occurs when parasites excreted by the parasitophorous vacuole. Many excellent reviews from leaders
the triatomine insect vector contaminate the bite wound or a mu- in the field have been published on the mechanistic aspects under-
cous membrane. In non-endemic areas, transmission may occur lying cell invasion (Alves and Colli, 2008, 2007; Yoshida and Cortez,
congenitally, via blood transfusion or organ transplantation, or as 2008; Mott and Burleigh, 2008; Mortara et al., 2005; Villalta et al.,
a result of a laboratory accident (Gutierrez et al., 2009). Despite a 2009).
century of scientific study, the relationship between the cell biol-
ogy of the host–parasite relationship and the pathophysiology of
Chagas disease remains incompletely understood. This review will 2. Tissue tropism

Infective metacyclic trypomastigotes inoculated into a wound


* Corresponding author. Address: Department of Pediatrics, Northwestern Uni-
versity, 320 Superior Street, Searle 4-409, Chicago, Illinois 60611, United States.
generally infect local macrophages, fibroblasts, and other mesen-
Fax: +1 312 503 1181. chymal tissues at the site of primary infection, followed by hema-
E-mail address: c-epting@northwestern.edu (C.L. Epting). togenous dissemination and stable infection of distant tissues

0014-4894/$ - see front matter Ó 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.exppara.2010.06.023
284 C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291

(Monteon et al., 1996) (Fig. 1). Although the parasite is capable of McGrath et al., 1995), play an important role in cellular infection,
infecting nearly any nucleated cells in vitro, a restricted tissue pool, and almost certainly are fundamental to permit passage through
involving cardiac and skeletal muscle and enteric nerves, develop the intact endothelium as well as the extracellular matrix. It has
apparent pathology (Zhang and Tarleton, 1999). It is tempting to also been recognized that surface residue modifications through
conclude that the parasite has intrinsic tissue tropism and, indeed, trans-sialidase contribute to endothelial cell interactions (Dias
this idea was initially established in classic work from Melo and et al., 2008). Further studies are needed to specifically address this
Brener (1978). Chagas disease demonstrates geographically-re- fundamental step in parasite dissemination through escape from
stricted clinical profiles (Macedo et al., 2004), lending support to the vascular compartment.
the notion of strain-dependent tissue-specific tropism, and genet-
ically distinct strains and clones can be isolated from patients with
primary cardiac or gastrointestinal disease (Vago et al., 2000). Fur- 3. Cell invasion
ther support for tissue tropism comes from the results of experi-
mental infection employing two isolates of the parasite, in which Interactions with host cells and the extracellular matrix occur
one strain was found to preferentially localize to the heart, and through a large and diverse group of surface glycoproteins and pro-
the other to the gastrointestinal tract (Andrade et al., 1999). How- teases. Since the pioneering work of (Dvorak and Hyde, 1973; Hyde
ever, the rich genetic variation in parasite population clearly con- and Dvorak, 1973), researchers have gained tremendous insight
tributes to disease outcome (Manoel-Caetano Fda and Silva, into the specific molecules involved during initial cell–cell interac-
2007), as does the host genetic background (Marinho et al., 2004; tions. Interestingly, many of the glycoproteins share the glycosyl-
Freitas et al., 2009). A clear molecular or immunologic explanation phosphatidylinositol (GPI) moiety. GPI-anchored proteins are first
for apparent tissue tropism is lacking and, at best, the hallmarks of synthesised in the ER, conjugated to a GPI-anchor, and attached
clinical disease appear to result from a complex interplay between to the membrane as luminal facing proteins in the endoplasmic
parasite and host genetic variation, inflammation, and immunity. reticulum. In the Golgi, they undergo extensive sugar and side-
Interestingly, and somewhat surprisingly, given the extensive chain modifications, and then fuse with the plasma membrane
literature on interactions between host leukocytes and vascular resulting in extracellular membrane-associated proteins (Brown
endothelium, the method of parasite egress from the bloodstream and Waneck, 1992; Tiede et al., 1999). The structures and functions
into the tissues remains to be established. The parasite is capable of these proteins are incredibly diverse, from adhesion, paracrine
of directly infecting the endothelium, and cardiac-specific studies signaling, surface enzymes, and cell differentiation (DosReis
of established Chagas heart disease demonstrate endothelial in- et al., 2002; Fujita and Jigami, 2008; Zacks and Garg, 2006). The
jury, inflammation, and microcirculatory compromise (Factor GPI-anchor confers several additional properties. First, enzymatic
et al., 1985; Tanowitz et al., 1996). The time-course for tissue dis- cleavage via glycosylphosphatidylinositol-specific phospholipase
semination is 7–10 days following inoculation (Monteon et al., C (GPI-PLC) can release the head group, and is implicated in lipid
1996), yet it is has not been established that parasites must first in- and paracrine signaling, as well as signal termination (Villalta
fect and lyse the vascular endothelium prior to spreading in the et al., 1998; Gaulton and Pratt, 1994). Additionally, GPI-anchored
surrounding tissues. Alternatively, the parasite could engage in proteins are thought to ubiquitously associate with, and in fact
regulated transmigration/diapedesis, stimulate a cytotoxic or may help define, the lipid raft microdomain compartment (Ken-
inflammatory injury resulting in breach of the permeability barrier, worthy et al., 2004) in other eukaryotic systems. Trypanosomes
enter and exit the endothelial cell without establishing an infec- were recognized early as cells with abundant expression of GPI-an-
tion, or escape via infected inflammatory cells acting as a Trojan chored proteins (Field et al., 1991), and these proteins form the
horse. The rich complement of surface proteases suggest that enzy- classic VSG coat critical to immune evasion by Trypanosoma brucei
matic digestion between the endothelial cell and into the underly- (Ferguson et al., 1994; Ferguson and Cross, 1984). Many GPI-an-
ing connective tissues is likely a direct, parasite-driven processes. chored proteins of T. cruzi are involved in both the host response
Indeed surface proteases, notably cruzipain (Stoka et al., 1995; and macrophage infection, as reviewed in Ropert et al. (2002). Gi-
ven the differences in synthesis and side-chain modifications, the
responsible enzymes are potential drug targets in the mammalian
host.
The mechanisms and route of cell invasion vary greatly with the
host cell type, and the reader is cautioned against broad general-
izations across cell types. Unlike some infectious agents that rely
on uptake and escape from professional phagocytic cells, T. cruzi
trypomastigotes are capable of directly invading both professional
phagocytes and non-phagocytic cells. Among the professional
phagocytes, tissue resident macrophages are critical targets for
early infection (Munoz-Fernandez et al., 1992), where they initiate
both innate immunity and the systemic anti-parasite inflammatory
response through epitope processing and presentation. Profes-
sional phagocytes have long been recognized both as necessary
cellular targets and as a defense mechanisms for the host. Macro-
phages form the backbone of the infection models exploring the
professional phagocytes–parasite interaction. These cells success-
fully harbor infection (Tanowitz et al., 1975) yet limit their own
infection (McCabe et al., 1984), likely through oxidative burst-
dependent killing (Bogdan and Rollinghoff, 1999), thus serving as
important parasite reservoirs. Trypomastigotes specifically induce
Fig. 1. Trypanosoma cruzi forms nests of intracellular parasites (white arrow) when
it infects mammalian cells, especially cardiac and skeletal muscle. Shown here is an
their uptake by professional phagocytes by engaging both TLR2
H&E stained section of a mouse heart demonstrating parasitosis of adjacent cardiac (Maganto-Garcia et al., 2008) and TLR 9-dependent pathways.
myocytes. Bar = 20 lm. The cellular mechanisms of phagocytosis have been well studied,
C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291 285

and several reviews are suggested (Mauel, 1982; Thorne and Black- of these surface protein families cannot be overemphasized. De-
well, 1983; De Araujo-Jorge et al., 1992; Kuhn, 1994). spite decades of study, a unified single invasion mechanism has
For infection of non-phagocytic cells, at least two major path- not emerged. Rather, a series of redundant and overlapping mech-
ways have been characterized. The first relies upon a calcium-med- anisms, varying with the parasite–host strain–strain combination,
iated signaling at the surface for lysosomal trafficking to provide have been reported (Alves and Colli, 2008). This diversity likely
donor membranes for the vacuole in a manner dependent upon ac- contributes to the co-evolutionary success of this parasite in the
tin polymerization and microtubules (Tardieux et al., 1994, 1992; infection of vertebrate cells and tissues.
Tyler and Engman, 2001; Schenkman et al., 1991), while the second The importance of the plasma membrane lipid environment is
is a plasma membrane-mediated invagination involving PI3 kinase rapidly gaining attention. Specialized regions, the lipid microdo-
signaling and independent of actin polymerization (Woolsey et al., mains/rafts, coordinate and regulate signaling events through tem-
2003; Burleigh, 2005; De Souza, 2002; Andrade and Andrews, poral-spatial organization of proteins. The kinetoplastids are no
2005). While these observations form a core understanding of cell exception, and GPI-anchored proteins are known to cluster in lipid
invasion, significant diversity, complexity, and redundancy in the rafts in this family (Denny et al., 2001). The host–parasite signaling
process have emerged over the past two decades. event likely depends upon surface–surface events coordinated
The capacity for cell invasion is not restricted to metacyclic or through lipid rafts, and indeed, cholesterol scavengers, which im-
cell-derived trypomastigotes. Both the dividing amastigotes (Mort- pair membrane fluidity and raft lateral reorganization, also impair
ara et al., 1999) and insect stage epimastigotes (Florencio-Martinez cell invasion (Fernandes et al., 2007; Barrias et al., 2007).
et al., 2010) are fully capable of establishing infections, and
amastigotes are increasingly recognized to share comparable infec-
tivity to trypomastigotes. Amastigote cell entry may follow more 5. Intracellular signaling and calcium
stereotyped pattern of cell invasion than the relatively diverse
patterns noted between trypomastigotes and their targets. Unlike The end result of host–parasite surface interaction is triggering
trypomastigotes, amastigotes invade in a Ca2+-dependent manner of bidirectional (host and parasite) signaling cascades which initi-
insensitive to PI3 inhibition but involving both cAMP signaling ate the invasion event. After extracellular matrix proteolysis and
and Ca2+-release from parasite acidocalcisomes. This has been surface binding through robust and redundant mechanisms, the
modeled in HeLa cells (Fernandes et al., 2006) and MDCK cells, parasite initiates a bidirectional calcium signaling cascade. This
where infection with G-strain amastigotes demonstrate a role for event can also be trigged in a cell-free system by isolated mem-
Rac1-mediated cell invasion (Fernandes and Mortara, 2004). While brane components or parasite lysate (Tardieux et al., 1994; Rodri-
this review will largely focus on pathways established in trypom- guez et al., 1995). This calcium signaling is fundamental to the
astigotes, do not discount the important role for infective amastig- downstream signaling cascade, which ends with the parasite en-
otes in propagating the local spread of infection within tissues of cased in an acidic parasitophorous vacuole (Mortara et al., 2005;
the parasitized host. Burleigh and Andrews, 1998). Some of the major signaling events
downstream of these surface receptors in the host and parasite try-
pomastigotes are tabulated (Table 2).
4. Surface interactions The precise molecular mechanisms leading to host and parasite
intracellular calcium release remain unknown. In the parasite, at
At the outset, parasites must survive, gain access to the cell sur- least two pathways have been identified. In metacyclic trypom-
face, and form stable attachments to host cells prior to entry. A astigotes, the engagement of gp82 with an unknown ligand trig-
cadre of protease-resistant surface glycoproteins either attach to gers a cascade in the parasite involving tyrosine phosphorylation
matrix components, bind cell surface receptors, or possess proteo- of p175 (Favoreto et al., 1998), the serine protein kinase C, and
lytic activity against matrix components. Many of these surface IP3-medated release of ER calcium stores (Yoshida et al., 2000).
molecules serve as adhesion anchors, some enable matrix destruc- An alternative, overlapping pathway occurs upon gp30 activation
tion or ligand cleavage, others help with immune evasion, and oth- (Cortez et al., 2003). Another major pathway, mediated by gp35/
ers initiate bidirectional signaling events in the parasite and host 50 binding to an unknown ligand, induces calcium release from
cell. Nearly 50% of the T. cruzi genome is dedicated to encoding acidocalcisomes through adenylate cyclase and a rise in cAMP
these surface proteins, broadly divided into several families: the (Yoshida and Cortez, 2008; Mortara et al., 2005), a pathway shared
gp63 surface proteases, the gp85/trans-sialidase superfamily (TS), during amastigote invasion. The protein tyrosine phosphatase
the mucins, and the mucin-associated surface proteins (Alves and gp90 is a negative regulator of invasion (Villalta et al., 2009; Todo-
Colli, 2008; Kulkarni et al., 2009; El-Sayed et al., 2005). A few of rov et al., 2000; Vieira et al., 2002; Zhong et al., 1998). Additionally,
the prominent surface glycoproteins and, if known, their ligands, TGF-b and integrin signaling on host cells have been implicated in
together with selected (not comprehensive) references are shown the invasion process, as have toll-like receptors (TLR2 and 9) (Cam-
(Table 1). A seminal example is the role of sialic acid in parasite vir- pos and Gazzinelli, 2004; Ming et al., 1995; Fernandez et al., 1993),
ulence (de Titto and Araujo, 1987). Surface mucins, a subset of the and the nerve growth factor receptor TrkA has been identified to
GPI-linked proteins, are modified by sialic acid scavenged from the bind to a trans-sialidase (Villalta et al., 2009). Signaling in the host
host through the action of the trans-sialidase (Jacobs et al., 2010), cell is even less well characterized. The generation of kinins by cru-
as the parasite lacks the ability to produce these modifications di- zipain results in bradykinin receptor (B2R)-mediated signaling
rectly. These sugar-modified residues, have been demonstrated to through PLC and IP3-kinase to release ER-bound calcium, opposed
have critical roles in cell attachment, invasion, and replication by the actions of the kininases (angiotensin converting enzyme-
(Muia et al., 2010). Other surface sugar residues upon glycopro- ACE) (Villalta et al., 2009; Scharfstein et al., 2000, 2008). Evidence
teins, notably mannose and galactose, also figure prominently in suggests that the anti-inflammatory properties of ACE inhibition is
the interaction and infection of host cells (Snary, 1985; Villalta useful to modulate cardiac inflammation in Chagas (Scharfstein
and Kierszenbaum, 1984, 1983). Many of the surface glycoproteins et al., 2008; Rassi et al., 2000), as it does in models of experimental
impact invasion or serve as virulence factors, since deletion, dis- autoimmune myocarditis (Daniels et al., 2007). Surface signaling
ruption of their enzymatic activity, or blockade of the receptor–li- through other bradykinin receptors (B1R) by the actions of kininase
gand interaction, usually reduces cell invasion in vitro and I, support invasion (Todorov et al., 2003), and the action of oligo-
improves outcome of infection in vivo. The extent and diversity peptidase B on its substrate is thought to generate an agonist for
286 C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291

Table 1
Surface glycoproteins of T. cruzi with extracellular matrix binding or proteolytic activity.

Protein Host target or ligand Biology Reference(s)


gp82 Mucin, unknown surface ligand Binding, signaling Neira et al. (2003)
gp63 Fibronectin, laminin ECM protease, binding Ortega-Barria and Pereira (1991)
Penetrin (gp60) Heparan, heparan sulfate, collagen Binding Ortega-Barria and Pereira (1991)
Tc-85/gp85/TS Fibronectin, laminin, cytokeratin 18 Binding, retention Colli (1993) and Katzin and Colli (1983)
gp35/50 Mucins Binding, signaling Ruiz et al. (1998)
gp90 Unknown Inhibitor of invasion, signaling Yoshida et al. (1990)
gp30 Unknown Binding Cortez et al. (2003)
Mucins/trans-sialidase 2,3-Sialyl containing host surface Sialidase, secreted (SAPA) immunogen Yoshida (2006) and Moody et al. (2000)
glycoproteins (galectin-3)
Mucin p45 Unknown Cardiac myocyte binding Turner et al. (2002)
gp83 Unknown Sialidase, Ca signaling Villalta et al. (2001)
Cruzipain Bradykinin Cysteine proteases Schenkman et al. (1991) and Souto-Padron et al. (1990)
POP Tc80 serine protease Collagen I, IV, fibronectin ECM protease Murta et al. (1990) and Santana et al. (1997)

Table 2
Signaling properties of selected T. cruzi surface glycoproteins.

Protein Biology Reference


gp82 Parasite: Ca increase, PLC dependent tyrosine phosphorylation of Tc-p175 Yoshida (2006) and Ramirez et al. (1993)
gp83 Parasite: Ca increase host: MAPK signaling Villalta et al. (1998, 1999)
gp30 Parasite: Ca increase Cortez et al. (2003)
gp35/50 Host and parasite cAMP and calcium increase Dorta et al. (1995), Yoshida et al. (1989) and Neira et al. (2002)
gp90 Phosphatase, potentially downregulates gp82 signaling Manque et al. (2003)
Cruzipain Bradykinin signaling, calcium increase, kinin generation Scharfstein et al. (2000)
Oligopeptidase B Cytosolic, cleaves a 120 kDa substrate, secreted, direct calcium release Caler et al. (1998) and Burleigh et al. (1997)

host cell calcium release through adenylate cyclase and phospholi- of the plasma membrane at the site of attachment in a wortman-
pase C (Caler et al., 1998). Additional receptors are proposed for li- nin-insensitive and lysosome-independent (Andrade and Andrews,
gand interactions with TS/Tc85, as well as additional substrates for 2004) manner. However, lysosomal fusion is thought to remain
cruzipain and chagasin, which interface with downstream signal- fundamental for a productive infection to occur through vacuole
ing in both the lysosome-dependent and independent pathways. acidification (Andrade and Andrews, 2004), and thus these diverse
Scharfstein and Lima recently published a detailed review on the results represent further insight into the components of the matur-
subject of the cysteine proteases, cruzipain, and protease inhibitors ing parasitophorous vacuole. While the precise molecular events
(Scharfstein and Lima, 2008). MAPK pathways have also been that lead to successful invasion have yet to be elucidated, the over-
implicated in macrophages through gp83 signaling (Villalta et al., arching theme is one of parasite entry through surface-initiated
1998). Alternative pathways for amastigote involving calcium re- signaling leading to a bidirectional rise in intracellular calcium,
lease from acidocalcisomes in a PI3 insensitive manner have al- causing reorganization, trafficking, and fusion of selected donor
ready been described above (Fernandes et al., 2006). membranes along the host cytoskeleton to the site of membrane
attachment and invagination.
6. Host membranes and the parasite vacuole
7. Host cytoskeleton
The classic model for parasite entry is based upon the rapid
recruitment of lysosomes to the parasite attachment point (An- The host cytoskeleton is critical for successful invasion. Host
drews, 1995) in a manner dependent upon microtubules and kine- cells are encased in an actin corset parallel to the inner membrane.
sin motors (Rodriquez et al., 1996). Host cell plasma membrane Calcium-mediated actin de-polymerization likely facilitates initial
and lysosomes have been assumed to be the donor membranes parasite entry and negatively impacts parasite retention (Tardieux
necessary for vacuole formation, and inhibition of membrane fu- et al., 1992; Woolsey et al., 2003; Woolsey and Burleigh, 2004). The
sion, vesicle trafficking, microtubule reorganization, molecular specific role of actin polymerization appears to depend on the spe-
motors, or calcium/cAMP signaling impairs successful invasion. cific cell type and parasite stage examined, with cytochalasin D
This vesicle-dependent pathway has been shown to be sensitive treatment enhancing trypomastigote, yet impairing amastigote,
to wortmannin, a PI3 kinase inhibitor, known to involve G-protein invasion (Procopio et al., 1998). A host of actin-associated elements
coupled receptors, and depend upon synaptotagmin-VII (Woolsey have been identified, including intermediate filaments, myosin-
et al., 2003; Leite et al., 1998; Caler et al., 2001). The precise char- associated components, integrins, and extracellular matrix compo-
acterization and sources of these donor membranes have become nents, as noted in a recent review (Mortara et al., 2008). The Rho/
more diverse with further investigation, including early and late Rac family of small GTPases is known to be a critical link between
endosomes (Woolsey et al., 2003; Wilkowsky et al., 2002), involve- surface signaling and changes in the underlying cytoskeleton.
ment of dynamin and Rab5 (Wilkowsky et al., 2002; Barrias et al., However, evidence suggests that trypomastigotes do not rely on
2010), and, recently, the autophagocytic pathway (Romano et al., this family for productive infection. In contrast, the invasion mech-
2009). Localized alterations in calcium concentration are known anism employed by amastigotes does depend upon Rac1 signaling,
signals for both microtubule-dependent lysosomal trafficking and again highlighting the diversity of cell invasion (Fernandes and
fusion (Luzio et al., 2007, 2003). More recently, this classic path- Mortara, 2004; Jou and Nelson, 1998). Members of the Rab family
way was usurped by a dominant alternative, a direct invagination of GTPases, necessary for endosomal compartment trafficking, are
C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291 287

essential for infection (Wilkowsky et al., 2002). The many compo- epimastigote differentiates into the metacyclic trypomastigote in
nents of the endosomal compartment (early, late, lysosomal, and the insect hindgut and, once introduced into the host, differenti-
autophagocytic) traffic along microtubules, which are necessary ates into the replicative intracellular amastigote. These ultimately
for infection. Evidence suggests that parasite entry itself may serve differentiate again into bloodform trypomastigotes, which lyse out
as a nucleation point for microtubule radiation from the parasi- of the cell. Both the amastigote and trypomastigote forms are capa-
tophorous vacuole membrane, further facilitating endosomal ble of propagating the local and metastatic infection. Transient
attraction initiated by calcium flux (Tyler et al., 2005). The relation- intermediate forms are thought to develop during the differentia-
ship between apparent parasite microtubule nucleation and lyso- tion process from amastigotes to trypomastigotes and appear in
somal attraction is unknown, nor is it understood if the parasite the mammalian host with the general morphology of epimastig-
stimulates this organization, or if this represents part of the host otes. (Tyler and Engman, 2001; Tonelli et al., 2004). Of specific
response to invasion. It remains possible that the forming vacuole interest here are the factors regulating intracellular differentiation.
is somehow attractive for c-tubulin, or that the vacuole simply Most notably is the acidic pH achieved in the vacuole. This environ-
‘‘sticks” to microtubules in the vicinity and that lysosome fusion ment initiates the differentiation program into the amastigote over
is a relatively passive rather than an active process. a period of several (2–8) h. In vitro the parasite spontaneously
undergoes differentiation if placed in an acidic environment (Toml-
inson et al., 1995). The replication of amastigotes also demon-
8. Cytoplasmic entry and parasite differentiation strates an absolute requirement for L-proline (Tonelli et al.,
2004), and the activity of phosphatases may be required for differ-
Now encased in the acidic parasitophorous vacuole (Fig. 2), par- entiation as well (Grellier et al., 1999). After a period of quiescence,
asite protection is offered by surface trans-sialidases (Hall et al., the amastigotes re-enter the cell cycle and undergo nine rounds of
1992), which also serve to facilitate parasite maturation and re- replication prior to further differentiation (or perhaps considered
lease (Rubin-de-Celis et al., 2006). There is clear evidence that de-differentiation) into motile trypomastigotes (Alves and Colli,
the parasite breaks down the vacuolar membrane to facilitate cyto- 2007; Andrade and Andrews, 2005). Interestingly, this process of
plasmic entry (de Carvalho and de Souza, 1989). The mechanism invasion, infection, and replication will occur even in cells stripped
for vacuolar escape is known to be lysosome and pH dependent of their nuclei, indicating that new host gene transcription is not
(Andrews, 1994). Early reports indicated that secretion of a por- necessary (Coimbra et al., 2007). The trypomastigotes destroy the
in-like/complement 9-related factor TcTOX (Andrews et al., host cell by unclear mechanisms, although evidence does not sup-
1990), and later a lytic factor LYT1 (Manning-Cela et al., 2001, port apoptotic cell death (Clark and Kuhn, 1999). After cytolysis,
2002) were critical for this final step in cell invasion. While TcTOX the infection cycle begins again for new host cell targets or uptake
has defined further molecular characterization, LYT1 null mutants by a naïve triatomine taking a bloodmeal.
(Manning-Cela et al., 2001) demonstrate markedly attenuated The transcriptional events in the trypanosome that regulate
infectivity (Zago et al., 2008). At the acidic pH of the vacuole these critical differentiation steps are poorly understood, and cur-
achieved through lysosomal fusion, LYT1 and/or TcTOX, are ex- rent dogma dictates that virtually all regulation occurs after tran-
pressed and assume conformations capable of promoting mem- scription and trans-splicing. Thus message stability, transcription
brane lysis to permit cytoplasmic entry. The invasive initiation, and post-transcriptional processing are critical events
trypomastigote thus functions as a loaded weapon, and, teleologi- for the kinetoplastids, which generate polycistronic transcripts at
cally, has completed its task to achieve successful invasion. a relatively constant rate, and control gene dosage largely through
As a digenetic organism, T. cruzi follows a differentiation contin- genomic copy amplification (Teixeira and daRocha, 2003; Wil-
uum from insect vector to mammalian host and back again. The liams, 1985; Teixeira, 1998). Not surprising, proteosome activity
is known to be essential for degrading stage-specific proteins dur-
ing the cytoskeletal remodeling that occurs during the transforma-
tion from trypomastigotes to amastigotes (Gonzalez et al., 1996).
Several T. cruzi-specific proteases and other enzymes have been
identified in the differentiation event (Meirelles et al., 1992; Harth
et al., 1993), but the upstream signals remain largely unknown.
Notably, at the transcriptional level, evidence suggests down-regu-
lation of RNA polymerases I and II occurs upon differentiation from
proliferative to non-proliferative forms (Elias et al., 2001). Addi-
tionally, stage-specific regulation of histone and ubiquitin genes
has been reported (Manning-Cela et al., 2006; Recinos et al.,
2001). Overall, identifying and understanding changes in gene
transcription, splicing, mRNA stability, and translational events
governing differentiation and replication are incompletely under-
stood and will benefit from additional study to develop specific
therapies targeting against T. cruzi.

9. Host response

Successful intracellular pathogens often co-opt the very cellular


self-preservation mechanisms designed to thwart parasitism. T.
cruzi has developed mechanisms of evading the immune responses
Fig. 2. T. cruzi interacts with the host cytoskeleton during cell invasion. Shown here and suppressing host apoptosis by modulating the expression of
is a H9C2 rat cardiomyoblast during invasion with Trypanosoma cruzi. Host and
parasite DNA in blue (DAPI) host lysosomes in red, host a-tubulin in green.
host cell surface receptors, secreted factors, and signaling mole-
Bar = 5 lm. (For interpretation of the references in colour in this figure legend, the cules. The pathogenesis of Chagas disease, and the relative contri-
reader is referred to the web version of this article.) butions of the parasite, inflammation, and autoimmunity remains a
288 C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291

matter of much investigation and debate, and is beyond the scope Andrade, L.O., Andrews, N.W., 2005. The Trypanosoma cruzi-host-cell interplay:
location, invasion, retention. Nat. Rev. Microbiol. 3, 819–823.
of this review.
Andrade, L.O., Machado, C.R., Chiari, E., Pena, S.D., Macedo, A.M., 1999. Differential
In addition to protective, anti-parasite immunity, individuals tissue distribution of diverse clones of Trypanosoma cruzi in infected mice. Mol.
with Chagas disease develop aberrant, potentially deleterious im- Biochem. Parasitol. 100, 163–172.
mune responses. Notably, invasion with the parasite triggers a type Andrews, N.W., 1994. From lysosomes into the cytosol: the intracellular pathway of
Trypanosoma cruzi. Braz. J. Med. Biol. Res. 27, 471–475.
I interferon response, known to be critical in during intracellular Andrews, N.W., 1995. Lysosome recruitment during host cell invasion by
invasion from bacteria and virus (a foreign protein/DNA/RNA re- Trypanosoma cruzi. Trends Cell Biol. 5, 133–137.
sponse), which may drive a local immune and autoimmune re- Andrews, N.W., Abrams, C.K., Slatin, S.L., Griffiths, G., 1990. A T. cruzi-secreted
protein immunologically related to the complement component C9: evidence
sponse (Chessler et al., 2008). An ancient cellular response to for membrane pore-forming activity at low pH. Cell 61, 1277–1287.
infection, termed the IFN-stimulatory DNA response (ISD), may Barrias, E.S., Dutra, J.M., De Souza, W., Carvalho, T.M., 2007. Participation of
be a key player in the local and adaptive immune response (Chess- macrophage membrane rafts in Trypanosoma cruzi invasion process. Biochem.
Biophys. Res. Commun. 363, 828–834.
ler et al., 2009). Both TLR2 and TLR9 mediated innate immune re- Barrias, E.S., Reignault, L.C., De Souza, W., Carvalho, T.M., 2010. Dynasore, a dynamin
sponse from presentation of T. cruzi methylated CpG antigens inhibitor, inhibits Trypanosoma cruzi entry into peritoneal macrophages. PLoS
(Bartholomeu et al., 2008) are involved (reviewed in Tarleton One 5, e7764.
Bartholomeu, D.C., Ropert, C., Melo, M.B., Parroche, P., Junqueira, C.F., Teixeira, S.M.,
(2007). Detailed reviews of the TLR-dependent and independent Sirois, C., Kasperkovitz, P., Knetter, C.F., Lien, E., Latz, E., Golenbock, D.T.,
pathways are noteworthy (Kayama and Takeda, 2010), as well as Gazzinelli, R.T., 2008. Recruitment and endo-lysosomal activation of TLR9 in
the role of innate immunity in clinical disease persistence (Sath- dendritic cells infected with Trypanosoma cruzi. J. Immunol. 181, 1333–1344.
Bogdan, C., Rollinghoff, M., 1999. How do protozoan parasites survive inside
ler-Avelar et al., 2009). In experimental models, dysregulation of
macrophages? Parasitol. Today 15, 22–28.
several components of the type I interferon response, notably the Brown, D., Waneck, G.L., 1992. Glycosyl-phosphatidylinositol-anchored membrane
members of the IFN regulatory factors (1–4) and their cognate proteins. J. Am. Soc. Nephrol. 3, 895–906.
binding proteins, drive spontaneous autoimmunity (Taki, 2002). Burleigh, B.A., 2005. Host cell signaling and Trypanosoma cruzi invasion: do all roads
lead to lysosomes? Sci. STKE pe36.
Stetson et al. found that deletion of the negative regulator Trex1, Burleigh, B.A., Andrews, N.W., 1998. Signaling and host cell invasion by
which serves to downregulate the ISD response, resulting in spon- Trypanosoma cruzi. Curr. Opin. Microbiol. 1, 461–465.
taneous cardiac autoimmunity (Stetson et al., 2008). It is intriguing Burleigh, B.A., Caler, E.V., Webster, P., Andrews, N.W., 1997. A cytosolic serine
endopeptidase from Trypanosoma cruzi is required for the generation of Ca2+
to surmise that the cardiac autoimmunity observed in Chagas is signaling in mammalian cells. J. Cell Biol. 136, 609–620.
the product of an imbalanced IFN response or failure to reset the Caler, E.V., Vaena de Avalos, S., Haynes, P.A., Andrews, N.W., Burleigh, B.A., 1998.
inflammatory response even after parasite clearance. Further study Oligopeptidase B-dependent signaling mediates host cell invasion by
Trypanosoma cruzi. EMBO J. 17, 4975–4986.
will be necessary to understand the involvement of the ISD re- Caler, E.V., Chakrabarti, S., Fowler, K.T., Rao, S., Andrews, N.W., 2001. The
sponse as a driver of both adaptive immunity and the resultant car- Exocytosis-regulatory protein synaptotagmin VII mediates cell invasion by
diac autoimmunity. Trypanosoma cruzi. J. Exp. Med. 193, 1097–1104.
Campos, M.A., Gazzinelli, R.T., 2004. Trypanosoma cruzi and its components as
exogenous mediators of inflammation recognized through Toll-like receptors.
10. Perspective Mediat. Inflamm. 13, 139–143.
Chessler, A.D., Ferreira, L.R., Chang, T.H., Fitzgerald, K.A., Burleigh, B.A., 2008. A novel
IFN regulatory factor 3-dependent pathway activated by trypanosomes triggers
Understanding the cellular interaction between parasite and IFN-beta in macrophages and fibroblasts. J. Immunol. 181, 7917–7924.
host and the host immune response are fundamental to treating Chessler, A.D., Unnikrishnan, M., Bei, A.K., Daily, J.P., Burleigh, B.A., 2009.
Trypanosoma cruzi triggers an early type I IFN response in vivo at the site of
parasitosis and preventing progression to Chagas disease. Decades intradermal infection. J. Immunol. 182, 2288–2296.
of research have revealed an incredibly rich surface proteome that Clark, R.K., Kuhn, R.E., 1999. Trypanosoma cruzi does not induce apoptosis in murine
confers some degree of tissue specificity while retaining broad plas- fibroblasts. Parasitology 118 (Pt 2), 167–175.
Coimbra, V.C., Yamamoto, D., Khusal, K.G., Atayde, V.D., Fernandes, M.C., Mortara,
ticity for cellular invasion. Pathways characterizing the upstream
R.A., Yoshida, N., Alves, M.J., Rabinovitch, M., 2007. Enucleated L929 cells
and downstream signaling events mediating cell invasion are par- support invasion, differentiation, and multiplication of Trypanosoma cruzi
tially understood at best, and many critical steps lack ligand–recep- parasites. Infect. Immun. 75, 3700–3706.
Colli, W., 1993. Trans-sialidase: a unique enzyme activity discovered in the
tor pairing. With the developments in advanced proteomic and
protozoan Trypanosoma cruzi. FASEB J. 7, 1257–1264.
lipidomic analysis, the time is right to dissect the surface interac- Cortez, M., Neira, I., Ferreira, D., Luquetti, A.O., Rassi, A., Atayde, V.D., Yoshida, N.,
tions between host and parasite. Our understanding of parasite dif- 2003. Infection by Trypanosoma cruzi metacyclic forms deficient in gp82 but
ferentiation, from trypomastigote to amastigote and back, remains expressing a related surface molecule, gp30. Infect. Immun. 71, 6184–6191.
Coura, J.R., 2007. Chagas disease: what is known and what is needed – a background
incomplete, and detailed investigation into parasite transcriptional article. Mem. Inst. Oswaldo Cruz 102 (Suppl. 1), 113–122.
and transcriptional regulation underlying cell differentiation is Coura, J.R., Dias, J.C., 2009. Epidemiology, control and surveillance of Chagas disease
likely to yield important insights. Finally, the emerging role of in- – 100 years after its discovery. Mem. Inst. Oswaldo Cruz 104, 31–40.
Daniels, M.D., Hyland, K.V., Engman, D.M., 2007. Treatment of experimental
nate cellular immunity both in facilitating effective cell invasion myocarditis via modulation of the renin–angiotensin system. Curr. Pharm.
and differentiation, and perhaps in affording resistance to over- Des. 13, 1299–1305.
whelming infection, is just now beginning to be investigated. De Araujo-Jorge, T.C., Barbosa, H.S., Meirelles, M.N., 1992. Trypanosoma cruzi
recognition by macrophages and muscle cells: perspectives after a 15-year
study. Mem. Inst. Oswaldo Cruz 87 (Suppl. 5), 43–56.
Acknowledgments de Carvalho, T.M., de Souza, W., 1989. Early events related with the behaviour of
Trypanosoma cruzi within an endocytic vacuole in mouse peritoneal
macrophages. Cell Struct. Funct. 14, 383–392.
We are grateful to Cheryl L Olson for the photomicrographs pro- De Souza, W., 2002. Basic cell biology of Trypanosoma cruzi. Curr. Pharm. Des. 8,
vided for this review. CLE and DME were supported in part by 269–285.
Grants from the United States National Institutes of Health. de Titto, E.H., Araujo, F.G., 1987. Mechanism of cell invasion by Trypanosoma cruzi:
importance of sialidase activity. Acta Trop. 44, 273–282.
Denny, P.W., Field, M.C., Smith, D.F., 2001. GPI-anchored proteins and
References glycoconjugates segregate into lipid rafts in Kinetoplastida. FEBS Lett. 491,
148–153.
Dias, W.B., Fajardo, F.D., Graca-Souza, A.V., Freire-de-Lima, L., Vieira, F., Girard, M.F.,
Alves, M.J., Colli, W., 2007. Trypanosoma cruzi: adhesion to the host cell and
Bouteille, B., Previato, J.O., Mendonca-Previato, L., Todeschini, A.R., 2008.
intracellular survival. IUBMB Life 59, 274–279.
Endothelial cell signalling induced by trans-sialidase from Trypanosoma cruzi.
Alves, M.J., Colli, W., 2008. Role of the gp85/trans-sialidase superfamily of
Cell. Microbiol. 10, 88–99.
glycoproteins in the interaction of Trypanosoma cruzi with host structures.
Dorta, M.L., Ferreira, A.T., Oshiro, M.E., Yoshida, N., 1995. Ca2+ signal induced by
Subcell. Biochem. 47, 58–69.
Trypanosoma cruzi metacyclic trypomastigote surface molecules implicated in
Andrade, L.O., Andrews, N.W., 2004. Lysosomal fusion is essential for the retention
mammalian cell invasion. Mol. Biochem. Parasitol. 73, 285–289.
of Trypanosoma cruzi inside host cells. J. Exp. Med. 200, 1135–1143.
C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291 289

DosReis, G.A., Pecanha, L.M., Bellio, M., Previato, J.O., Mendonca-Previato, L., 2002. Jacobs, T., Erdmann, H., Fleischer, B., 2010. Molecular interaction of Siglecs (sialic
Glycoinositol phospholipids from Trypanosoma cruzi transmit signals to the acid-binding Ig-like lectins) with sialylated ligands on Trypanosoma cruzi. Eur. J.
cells of the host immune system through both ceramide and glycan chains. Cell Biol. 89, 113–116.
Microbes Infect. 4, 1007–1013. Jou, T.S., Nelson, W.J., 1998. Effects of regulated expression of mutant RhoA and
Dvorak, J.A., Hyde, T.P., 1973. Trypanosoma cruzi: interaction with vertebrate cells Rac1 small GTPases on the development of epithelial (MDCK) cell polarity. J.
in vitro. I. Individual interactions at the cellular and subcellular levels. Exp. Cell Biol. 142, 85–100.
Parasitol. 34, 268–283. Katzin, A.M., Colli, W., 1983. Lectin receptors in Trypanosoma cruzi. An N-acetyl-D-
Elias, M.C., Marques-Porto, R., Freymuller, E., Schenkman, S., 2001. Transcription glucosamine-containing surface glycoprotein specific for the trypomastigote
rate modulation through the Trypanosoma cruzi life cycle occurs in parallel with stage. Biochim. Biophys. Acta 727, 403–411.
changes in nuclear organisation. Mol. Biochem. Parasitol. 112, 79–90. Kayama, H., Takeda, K., 2010. The innate immune response to Trypanosoma cruzi
El-Sayed, N.M., Myler, P.J., Bartholomeu, D.C., Nilsson, D., Aggarwal, G., Tran, A.N., infection. Microbes Infect. 12, 511–517.
Ghedin, E., Worthey, E.A., Delcher, A.L., Blandin, G., Westenberger, S.J., Caler, E., Kenworthy, A.K., Nichols, B.J., Remmert, C.L., Hendrix, G.M., Kumar, M., Zimmerberg,
Cerqueira, G.C., Branche, C., Haas, B., Anupama, A., Arner, E., Aslund, L., Attipoe, J., Lippincott-Schwartz, J., 2004. Dynamics of putative raft-associated proteins
P., Bontempi, E., Bringaud, F., Burton, P., Cadag, E., Campbell, D.A., Carrington, at the cell surface. J. Cell Biol. 165, 735–746.
M., Crabtree, J., Darban, H., da Silveira, J.F., de Jong, P., Edwards, K., Englund, P.T., Kuhn, R.E., 1994. Macrophages in experimental Chagas’ disease. Immunol. Ser. 60,
Fazelina, G., Feldblyum, T., Ferella, M., Frasch, A.C., Gull, K., Horn, D., Hou, L., 495–502.
Huang, Y., Kindlund, E., Klingbeil, M., Kluge, S., Koo, H., Lacerda, D., Levin, M.J., Kulkarni, M.M., Olson, C.L., Engman, D.M., McGwire, B.S., 2009. Trypanosoma cruzi
Lorenzi, H., Louie, T., Machado, C.R., McCulloch, R., McKenna, A., Mizuno, Y., GP63 proteins undergo stage-specific differential posttranslational modification
Mottram, J.C., Nelson, S., Ochaya, S., Osoegawa, K., Pai, G., Parsons, M., Pentony, and are important for host cell infection. Infect. Immun. 77, 2193–2200.
M., Pettersson, U., Pop, M., Ramirez, J.L., Rinta, J., Robertson, L., Salzberg, S.L., Leite, M.F., Moyer, M.S., Andrews, N.W., 1998. Expression of the mammalian
Sanchez, D.O., Seyler, A., Sharma, R., Shetty, J., Simpson, A.J., Sisk, E., Tammi, calcium signaling response to Trypanosoma cruzi in Xenopus laevis oocytes. Mol.
M.T., Tarleton, R., Teixeira, S., Van Aken, S., Vogt, C., Ward, P.N., Wickstead, B., Biochem. Parasitol. 92, 1–13.
Wortman, J., White, O., Fraser, C.M., Stuart, K.D., Andersson, B., 2005. The Luzio, J.P., Poupon, V., Lindsay, M.R., Mullock, B.M., Piper, R.C., Pryor, P.R., 2003.
genome sequence of Trypanosoma cruzi, etiologic agent of Chagas disease. Membrane dynamics and the biogenesis of lysosomes. Mol. Membr. Biol. 20,
Science 309, 409–415. 141–154.
Factor, S.M., Cho, S., Wittner, M., Tanowitz, H., 1985. Abnormalities of the coronary Luzio, J.P., Bright, N.A., Pryor, P.R., 2007. The role of calcium and other ions in sorting
microcirculation in acute murine Chagas’ disease. Am. J. Trop. Med. Hyg. 34, and delivery in the late endocytic pathway. Biochem. Soc. Trans. 35, 1088–1091.
246–253. Macedo, A.M., Machado, C.R., Oliveira, R.P., Pena, S.D., 2004. Trypanosoma cruzi:
Favoreto Jr., S., Dorta, M.L., Yoshida, N., 1998. Trypanosoma cruzi 175-kDa protein genetic structure of populations and relevance of genetic variability to the
tyrosine phosphorylation is associated with host cell invasion. Exp. Parasitol. pathogenesis of chagas disease. Mem. Inst. Oswaldo Cruz 99, 1–12.
89, 188–194. Maganto-Garcia, E., Punzon, C., Terhorst, C., Fresno, M., 2008. Rab5 activation by
Ferguson, M.A., Cross, G.M.A., 1984. Myristylation of the membrane form of a Toll-like receptor 2 is required for Trypanosoma cruzi internalization and
Trypanosoma brucei variant surface glycoprotein. J. Biol. Chem. 259, 3011–3015. replication in macrophages. Traffic 9, 1299–1315.
Ferguson, M.A., Brimacombe, J.S., Cottaz, S., Field, R.A., Guther, L.S., Homans, S.W., Manning-Cela, R., Cortes, A., Gonzalez-Rey, E., Van Voorhis, W.C., Swindle, J.,
McConville, M.J., Mehlert, A., Milne, K.G., Ralton, J.E., et al., 1994. Glycosyl- Gonzalez, A., 2001. LYT1 protein is required for efficient in vitro infection by
phosphatidylinositol molecules of the parasite and the host. Parasitology 108 Trypanosoma cruzi. Infect. Immun. 69, 3916–3923.
(Suppl.), S45–S54. Manning-Cela, R., Gonzalez, A., Swindle, J., 2002. Alternative splicing of LYT1
Fernandes, A.B., Mortara, R.A., 2004. Invasion of MDCK epithelial cells with altered transcripts in Trypanosoma cruzi. Infect. Immun. 70, 4726–4728.
expression of Rho GTPases by Trypanosoma cruzi amastigotes and metacyclic Manning-Cela, R., Jaishankar, S., Swindle, J., 2006. Life-cycle and growth-phase-
trypomastigotes of strains from the two major phylogenetic lineages. Microbes dependent regulation of the ubiquitin genes of Trypanosoma cruzi. Arch. Med.
Infect. 6, 460–467. Res. 37, 593–601.
Fernandes, A.B., Neira, I., Ferreira, A.T., Mortara, R.A., 2006. Cell invasion by Manoel-Caetano Fda, S., Silva, A.E., 2007. Implications of genetic variability of
Trypanosoma cruzi amastigotes of distinct infectivities: studies on signaling Trypanosoma cruzi for the pathogenesis of Chagas disease. Cad Saude Publica 23,
pathways. Parasitol. Res. 100, 59–68. 2263–2274.
Fernandes, M.C., Cortez, M., Geraldo Yoneyama, K.A., Straus, A.H., Yoshida, N., Manque, P.M., Neira, I., Atayde, V.D., Cordero, E., Ferreira, A.T., da Silveira, J.F.,
Mortara, R.A., 2007. Novel strategy in Trypanosoma cruzi cell invasion: Ramirez, M., Yoshida, N., 2003. Cell adhesion and Ca2+ signaling activity in
implication of cholesterol and host cell microdomains. Int. J. Parasitol. 37, stably transfected Trypanosoma cruzi epimastigotes expressing the metacyclic
1431–1441. stage-specific surface molecule gp82. Infect. Immun. 71, 1561–1565.
Fernandez, M.A., Munoz-Fernandez, M.A., Fresno, M., 1993. Involvement of beta 1 Marinho, C.R., Bucci, D.Z., Dagli, M.L., Bastos, K.R., Grisotto, M.G., Sardinha, L.R.,
integrins in the binding and entry of Trypanosoma cruzi into human Baptista, C.R., Goncalves, C.P., Lima, M.R., Alvarez, J.M., 2004. Pathology affects
macrophages. Eur. J. Immunol. 23, 552–557. different organs in two mouse strains chronically infected by a Trypanosoma
Field, M.C., Medina-Acosta, E., Cross, G.A., 1991. Characterization of a cruzi clone: a model for genetic studies of Chagas’ disease. Infect. Immun. 72,
glycosylphosphatidylinositol membrane protein anchor precursor in 2350–2357.
Leishmania mexicana. Mol. Biochem. Parasitol. 48, 227–229. Mauel, J., 1982. In vitro induction of intracellular killing of parasitic protozoa by
Florencio-Martinez, L., Marquez-Duenas, C., Ballesteros-Rodea, G., Martinez- macrophages. Immunobiology 161, 392–400.
Calvillo, S., Manning-Cela, R., 2010. Cellular analysis of host cell infection by McCabe, R.E., Remington, J.S., Araujo, F.G., 1984. Mechanisms of invasion and
different developmental stages of Trypanosoma cruzi. Exp. Parasitol. 126, replication of the intracellular stage in Trypanosoma cruzi. Infect. Immun. 46,
332–336. 372–376.
Freitas, J.M., Andrade, L.O., Pires, S.F., Lima, R., Chiari, E., Santos, R.R., Soares, M., McGrath, M.E., Eakin, A.E., Engel, J.C., McKerrow, J.H., Craik, C.S., Fletterick, R.J., 1995.
Machado, C.R., Franco, G.R., Pena, S.D., Macedo, A.M., 2009. The MHC gene The crystal structure of cruzain: a therapeutic target for Chagas’ disease. J. Mol.
region of murine hosts influences the differential tissue tropism of infecting Biol. 247, 251–259.
Trypanosoma cruzi strains. PLoS One 4, e5113. Meirelles, M.N., Juliano, L., Carmona, E., Silva, S.G., Costa, E.M., Murta, A.C.,
Fujita, M., Jigami, Y., 2008. Lipid remodeling of GPI-anchored proteins and its Scharfstein, J., 1992. Inhibitors of the major cysteinyl proteinase (GP57/51)
function. Biochim. Biophys. Acta 1780, 410–420. impair host cell invasion and arrest the intracellular development of
Gaulton, G.N., Pratt, J.C., 1994. Glycosylated phosphatidylinositol molecules as Trypanosoma cruzi in vitro. Mol. Biochem. Parasitol. 52, 175–184.
second messengers. Semin. Immunol. 6, 97–104. Melo, R.C., Brener, Z., 1978. Tissue tropism of different Trypanosoma cruzi strains. J.
Gonzalez, J., Ramalho-Pinto, F.J., Frevert, U., Ghiso, J., Tomlinson, S., Scharfstein, J., Parasitol. 64, 475–482.
Corey, E.J., Nussenzweig, V., 1996. Proteasome activity is required for the stage- Milei, J., Guerri-Guttenberg, R.A., Grana, D.R., Storino, R., 2009. Prognostic impact of
specific transformation of a protozoan parasite. J. Exp. Med. 184, 1909–1918. Chagas disease in the United States. Am. Heart J. 157, 22–29.
Grellier, P., Blum, J., Santana, J., Bylen, E., Mouray, E., Sinou, V., Teixeira, A.R., Ming, M., Ewen, M.E., Pereira, M.E.A., 1995. Trypanosome invasion of mammalian
Schrevel, J., 1999. Involvement of calyculin A-sensitive phosphatase(s) in the cells requires activation of the TGF beta signalling pathway. Cell 82, 287–296.
differentiation of Trypanosoma cruzi trypomastigotes to amastigotes. Mol. Moncayo, A., Silveira, A.C., 2009. Current epidemiological trends for Chagas disease
Biochem. Parasitol. 98, 239–252. in Latin America and future challenges in epidemiology, surveillance and health
Gutierrez, F.R., Guedes, P.M., Gazzinelli, R.T., Silva, J.S., 2009. The role of parasite policy. Mem. Inst. Oswaldo Cruz 104 (Suppl. 1), 17–30.
persistence in pathogenesis of Chagas heart disease. Parasite Immunol. 31, 673– Monteon, V.M., Furuzawa-Carballeda, J., Alejandre-Aguilar, R., Aranda-Fraustro, A.,
685. Rosales-Encina, J.L., Reyes, P.A., 1996. American trypanosomosis: in situ and
Hall, B.F., Webster, P., Ma, A.K., Joiner, K.A., Andrews, N.W., 1992. Desialylation of generalized features of parasitism and inflammation kinetics in a murine
lysosomal membrane glycoproteins by Trypanosoma cruzi: a role for the surface model. Exp. Parasitol. 83, 267–274.
neuraminidase in facilitating parasite entry into the host cell cytoplasm. J. Exp. Moody, T.N., Ochieng, J., Villalta, F., 2000. Novel mechanism that Trypanosoma cruzi
Med. 176, 313–325. uses to adhere to the extracellular matrix mediated by human galectin-3. FEBS
Harth, G., Andrews, N., Mills, A.A., Engel, J.C., Smith, R., McKerrow, J.H., 1993. Lett. 470, 305–308.
Peptide-fluoromethyl ketones arrest intracellular replication and intercellular Mortara, R.A., Procopio, D.O., Barros, H.C., Verbisck, N.V., Andreoli, W.K., Silva, R.B.,
transmission of Trypanosoma cruzi. Mol. Biochem. Parasitol. 58, 17–24. da Silva, S., 1999. Features of host cell invasion by different infective forms of
Hyde, T.P., Dvorak, J.A., 1973. Trypanosoma cruzi: interaction with vertebrate cells Trypanosoma cruzi. Mem. Inst. Oswaldo Cruz 94 (Suppl. 1), 135–137.
in vitro. 2. Quantitative analysis of the penetration phase. Exp. Parasitol. 34, Mortara, R.A., Andreoli, W.K., Taniwaki, N.N., Fernandes, A.B., Silva, C.V., Fernandes,
284–294. M.C., L’Abbate, C., Silva, S., 2005. Mammalian cell invasion and intracellular
290 C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291

trafficking by Trypanosoma cruzi infective forms. An. Acad. Bras. Cienc. 77, 77– Schenkman, S., Robbins, E.S., Nussenzweig, V., 1991. Attachment of Trypanosoma
94. cruzi to mammalian cells requires parasite energy, and invasion can be
Mortara, R.A., Andreoli, W.K., Fernandes, M.C., da Silva, C.V., Fernandes, A.B., independent of the target cell cytoskeleton. Infect. Immun. 59, 645–654.
L’Abbate, C., da Silva, S., 2008. Host cell actin remodeling in response to Schmunis, G.A., 2007. Epidemiology of Chagas disease in non-endemic countries:
Trypanosoma cruzi: trypomastigote versus amastigote entry. Subcell. Biochem. the role of international migration. Mem. Inst. Oswaldo Cruz 102 (Suppl. 1), 75–
47, 101–109. 85.
Mott, G.A., Burleigh, B.A., 2008. The role of host cell lysosomes in Trypanosoma cruzi Snary, D., 1985. Receptors and recognition mechanisms of Trypanosoma cruzi. Trans.
invasion. Subcell. Biochem. 47, 165–173. Roy. Soc. Trop. Med. Hyg. 79, 587–590.
Muia, R.P., Yu, H., Prescher, J.A., Hellman, U., Chen, X., Bertozzi, C.R., Campetella, O., Souto-Padron, T., Campetella, O.E., Cazzulo, J.J., de Souza, W., 1990. Cysteine
2010. Identification of glycoproteins targeted by Trypanosoma cruzi trans- proteinase in Trypanosoma cruzi: immunocytochemical localization and
sialidase, a virulence factor that disturbs lymphocyte glycosylation. involvement in parasite–host cell interaction. J. Cell Sci. 96, 485–490.
Glycobiology. 20, 833–842. Stetson, D.B., Ko, J.S., Heidmann, T., Medzhitov, R., 2008. Trex1 prevents cell-
Munoz-Fernandez, M.A., Fernandez, M.A., Fresno, M., 1992. Synergism between intrinsic initiation of autoimmunity. Cell 134, 587–598.
tumor necrosis factor-alpha and interferon-gamma on macrophage activation Stoka, V., Nycander, M., Lenarcic, B., Labriola, C., Cazzulo, J.J., Bjork, I., Turk, V., 1995.
for the killing of intracellular Trypanosoma cruzi through a nitric oxide- Inhibition of cruzipain, the major cysteine proteinase of the protozoan parasite,
dependent mechanism. Eur. J. Immunol. 22, 301–307. Trypanosoma cruzi, by proteinase inhibitors of the cystatin superfamily. FEBS
Murta, A.C., Persechini, P.M., Padron, T., de Souza, W., Guimaraes, J.A., Scharfstein, J., Lett. 370, 101–104.
1990. Structural and functional identification of GP57/51 antigen of Taki, S., 2002. Type I interferons and autoimmunity: lessons from the clinic and
Trypanosoma cruzi as a cysteine proteinase. Mol. Biochem. Parasitol. 43, 27–38. from IRF-2-deficient mice. Cytokine Growth Factor Rev. 13, 379–391.
Neira, I., Ferreira, A.T., Yoshida, N., 2002. Activation of distinct signal transduction Tanowitz, H., Wittner, M., Kress, Y., Bloom, B., 1975. Studies of in vitro infection by
pathways in Trypanosoma cruzi isolates with differential capacity to invade host Trypanosoma cruzi. I. Ultrastructural studies on the invasion of macrophages
cells. Int. J. Parasitol. 32, 405–414. and L-cells. Am. J. Trop. Med. Hyg. 24, 25–33.
Neira, I., Silva, F.A., Cortez, M., Yoshida, N., 2003. Involvement of Trypanosoma cruzi Tanowitz, H.B., Kaul, D.K., Chen, B., Morris, S.A., Factor, S.M., Weiss, L.M., Wittner, M.,
metacyclic trypomastigote surface molecule gp82 in adhesion to gastric mucin 1996. Compromised microcirculation in acute murine Trypanosoma cruzi
and invasion of epithelial cells. Infect. Immun. 71, 557–561. infection. J. Parasitol. 82, 124–130.
Ortega-Barria, E., Pereira, M.E., 1991. A novel T. cruzi heparin-binding protein Tardieux, I., Webster, P., Ravesloot, J., Boron, W., Lunn, J.A., Heuser, J.E., Andrews,
promotes fibroblast adhesion and penetration of engineered bacteria and N.W., 1992. Lysosome recruitment and fusion are early events required for
trypanosomes into mammalian cells. Cell 67, 411–421. trypanosome invasion of mammalian cells. Cell 71, 1117–1130.
Perez-Molina, J.A., Perez-Ayala, A., Moreno, S., Fernandez-Gonzalez, M.C., Zamora, J., Tardieux, I., Nathanson, M.H., Andrews, N.W., 1994. Role in host cell invasion of
Lopez-Velez, R., 2009. Use of benznidazole to treat chronic Chagas’ disease: a Trypanosoma cruzi-induced cytosolic-free Ca2+ transients. J. Exp. Med. 179,
systematic review with a meta-analysis. J. Antimicrob. Chemother. 64, 1139–1147. 1017–1022.
Procopio, D.O., da Silva, S., Cunningham, C.C., Mortara, R.A., 1998. Trypanosoma Tarleton, R.L., 2007. Immune system recognition of Trypanosoma cruzi. Curr. Opin.
cruzi: effect of protein kinase inhibitors and cytoskeletal protein organization Immunol. 19, 430–434.
and expression on host cell invasion by amastigotes and metacyclic Teixeira, S.M., 1998. Control of gene expression in Trypanosomatidae. Braz. J. Med.
trypomastigotes. Exp. Parasitol. 90, 1–13. Biol. Res. 31, 1503–1516.
Ramirez, M.I., Ruiz Rde, C., Araya, J.E., Da Silveira, J.F., Yoshida, N., 1993. Involvement Teixeira, S.M., daRocha, W.D., 2003. Control of gene expression and genetic
of the stage-specific 82-kilodalton adhesion molecule of Trypanosoma cruzi manipulation in the Trypanosomatidae. Genet. Mol. Res. 2, 148–158.
metacyclic trypomastigotes in host cell invasion. Infect. Immun. 61, 3636–3641. Thorne, K.J., Blackwell, J.M., 1983. Cell-mediated killing of protozoa. Adv. Parasitol.
Rassi Jr., A., Rassi, A., Little, W.C., 2000. Chagas’ heart disease. Clin. Cardiol. 23, 883– 22, 43–151.
889. Tiede, A., Bastisch, I., Schubert, J., Orlean, P., Schmidt, R.E., 1999. Biosynthesis of
Rassi Jr., A., Dias, J.C., Marin-Neto, J.A., Rassi, A., 2009. Challenges and opportunities glycosylphosphatidylinositols in mammals and unicellular microbes. Biol.
for primary, secondary, and tertiary prevention of Chagas’ disease. Heart 95, Chem. 380, 503–523.
524–534. Todorov, A.G., Einicker-Lamas, M., de Castro, S.L., Oliveira, M.M., Guilherme, A.,
Recinos, R.F., Kirchhoff, L.V., Donelson, J.E., 2001. Cell cycle expression of histone 2000. Activation of host cell phosphatidylinositol 3-kinases by Trypanosoma
genes in Trypanosoma cruzi. Mol. Biochem. Parasitol. 113, 215–222. cruzi infection. J. Biol. Chem. 275, 32182–32186.
Rodriguez, A., Rioult, M.G., Ora, A., Andrews, N.W., 1995. A trypanosome-soluble Todorov, A.G., Andrade, D., Pesquero, J.B., Araujo Rde, C., Bader, M., Stewart, J., Gera,
factor induces IP3 formation, intracellular Ca2+ mobilization and microfilament L., Muller-Esterl, W., Morandi, V., Goldenberg, R.C., Neto, H.C., Scharfstein, J.,
rearrangement in host cells. J. Cell Biol. 129, 1263–1273. 2003. Trypanosoma cruzi induces edematogenic responses in mice and invades
Rodriquez, A., Samoff, E., Rioult, M.G., Chung, A., Andrews, N.W., 1996. Host cell cardiomyocytes and endothelial cells in vitro by activating distinct kinin
invasion by trypanosomes requires lysosomes and microtubule/kinesin- receptor (B1/B2) subtypes. FASEB J. 17, 73–75.
mediated transport. J. Cell Biol. 134, 349–362. Tomlinson, S., Vandekerckhove, F., Frevert, U., Nussenzweig, V., 1995. The induction
Romano, P.S., Arboit, M.A., Vazquez, C.L., Colombo, M.I., 2009. The autophagic of Trypanosoma cruzi trypomastigote to amastigote transformation by low pH.
pathway is a key component in the lysosomal dependent entry of Trypanosoma Parasitology 110, 547–554.
cruzi into the host cell. Autophagy 5, 6–18. Tonelli, R.R., Silber, A.M., Almeida-de-Faria, M., Hirata, I.Y., Colli, W., Alves, M.J.,
Ropert, C., Ferreira, L.R., Campos, M.A., Procopio, D.O., Travassos, L.R., Ferguson, M.A., 2004. L-Proline is essential for the intracellular differentiation of Trypanosoma
Reis, L.F., Teixeira, M.M., Almeida, I.C., Gazzinelli, R.T., 2002. Macrophage cruzi. Cell. Microbiol. 6, 733–741.
signaling by glycosylphosphatidylinositol-anchored mucin-like glycoproteins Turner, C.W., Lima, M.F., Villalta, F., 2002. Trypanosoma cruzi uses a 45-kDa mucin
derived from Trypanosoma cruzi trypomastigotes. Microbes Infect. 4, 1015–1025. for adhesion to mammalian cells. Biochem. Biophys. Res. Commun. 290, 29–34.
Rubin-de-Celis, S.S., Uemura, H., Yoshida, N., Schenkman, S., 2006. Expression of Tyler, K.M., Engman, D.M., 2001. The life cycle of Trypanosoma cruzi revisited. Int. J.
trypomastigote trans-sialidase in metacyclic forms of Trypanosoma cruzi Parasitol. 31, 472–481.
increases parasite escape from its parasitophorous vacuole. Cell. Microbiol. 8, Tyler, K.M., Luxton, G.W., Applewhite, D.A., Murphy, S.C., Engman, D.M., 2005.
1888–1898. Responsive microtubule dynamics promote cell invasion by Trypanosoma cruzi.
Ruiz, R.C., Favoreto Jr., S., Dorta, M.L., Oshiro, M.E., Ferreira, A.T., Manque, P.M., Cell. Microbiol. 7, 1579–1591.
Yoshida, N., 1998. Infectivity of Trypanosoma cruzi strains is associated with Vago, A.R., Andrade, L.O., Leite, A.A., d’Avila Reis, D., Macedo, A.M., Adad, S.J., Tostes
differential expression of surface glycoproteins with differential Ca2+ signalling Jr., S., Moreira, M.C., Filho, G.B., Pena, S.D., 2000. Genetic characterization of
activity. Biochem. J. 330, 505–511. Trypanosoma cruzi directly from tissues of patients with chronic Chagas disease:
Santana, J.M., Grellier, P., Schrevel, J., Teixeira, A.R., 1997. A Trypanosoma cruzi- differential distribution of genetic types into diverse organs. Am. J. Pathol. 156,
secreted 80 kDa proteinase with specificity for human collagen types I and IV. 1805–1809.
Biochem. J. 325 (Pt 1), 129–137. Vieira, M., Dutra, J.M., Carvalho, T.M., Cunha-e-Silva, N.L., Souto-Padron, T., Souza,
Sathler-Avelar, R., Vitelli-Avelar, D.M., Teixeira-Carvalho, A., Martins-Filho, O.A., W., 2002. Cellular signaling during the macrophage invasion by Trypanosoma
2009. Innate immunity and regulatory T-cells in human Chagas disease: what cruzi. Histochem. Cell Biol. 118, 491–500.
must be understood? Mem. Inst. Oswaldo Cruz 104 (Suppl. 1), 246–251. Villalta, F., Kierszenbaum, F., 1983. Role of cell surface mannose residues in host cell
Scharfstein, J., Lima, A.P., 2008. Roles of naturally occurring protease inhibitors in invasion by Trypanosoma cruzi. Biochim. Biophys. Acta 736, 39–44.
the modulation of host cell signaling and cellular invasion by Trypanosoma cruzi. Villalta, F., Kierszenbaum, F., 1984. Host cell invasion by Trypanosoma cruzi: role of
Subcell. Biochem. 47, 140–154. cell surface galactose residues. Biochem. Biophys. Res. Commun. 119, 228–235.
Scharfstein, J., Schmitz, V., Morandi, V., Capella, M.M., Lima, A.P., Morrot, A., Juliano, Villalta, F., Zhang, Y., Bibb, K.E., Burns Jr., J.M., Lima, M.F., 1998. Signal transduction
L., Muller-Esterl, W., 2000. Host cell invasion by Trypanosoma cruzi is in human macrophages by gp83 ligand of Trypanosoma cruzi: trypomastigote
potentiated by activation of bradykinin B(2) receptors. J. Exp. Med. 192, gp83 ligand up-regulates trypanosome entry through the MAP kinase pathway.
1289–1300. Biochem. Biophys. Res. Commun. 249, 247–252.
Scharfstein, J., Monteiro, A.C., Schmitz, V., Svensjo, E., 2008. Angiotensin-converting Villalta, F., Zhang, Y., Bibb, K.E., Pratap, S., Burns Jr., J.M., Lima, M.F., 1999. Signal
enzyme limits inflammation elicited by Trypanosoma cruzi cysteine proteases: a transduction in human macrophages by gp83 ligand of Trypanosoma cruzi:
peripheral mechanism regulating adaptive immunity via the innate kinin trypomastigote gp83 ligand up-regulates trypanosome entry through protein
pathway. Biol. Chem. 389, 1015–1024. kinase C activation. Mol. Cell Biol. Res. Commun. 2, 64–70.
Schenkman, S., Diaz, C., Nussenzweig, V., 1991. Attachment of Trypanosoma cruzi Villalta, F., Smith, C.M., Ruiz-Ruano, A., Lima, M.F., 2001. A ligand that Trypanosoma
trypomastigotes to receptors at restricted cell surface domains. Exp. Parasitol. cruzi uses to bind to mammalian cells to initiate infection. FEBS Lett. 505, 383–
72, 76–86. 388.
C.L. Epting et al. / Experimental Parasitology 126 (2010) 283–291 291

Villalta, F., Scharfstein, J., Ashton, A.W., Tyler, K.M., Guan, F., Mukherjee, S., Lima, Yoshida, N., Mortara, R.A., Araguth, M.F., Gonzalez, J.C., Russo, M., 1989. Metacyclic
M.F., Alvarez, S., Weiss, L.M., Huang, H., Machado, F.S., Tanowitz, H.B., 2009. neutralizing effect of monoclonal antibody 10D8 directed to the 35- and 50-
Perspectives on the Trypanosoma cruzi-host cell receptor interactions. Parasitol. kilodalton surface glycoconjugates of Trypanosoma cruzi. Infect. Immun. 57,
Res. 104, 1251–1260. 1663–1667.
Wilkowsky, S.E., Barbieri, M.A., Stahl, P.D., Isola, E.L., 2002. Regulation of Yoshida, N., Blanco, S.A., Araguth, M.F., Russo, M., Gonzalez, J., 1990. The stage-
Trypanosoma cruzi invasion of nonphagocytic cells by the endocytically active specific 90-kilodalton surface antigen of metacyclic trypomastigotes of
GTPases dynamin, Rab5, and Rab7. Biochem. Biophys. Res. Commun. 291, 516– Trypanosoma cruzi. Mol. Biochem. Parasitol. 39, 39–46.
521. Yoshida, N., Favoreto Jr., S., Ferreira, A.T., Manque, P.M., 2000. Signal transduction
Williams, G.T., 1985. Control of differentiation in Trypanosoma cruzi. Curr. Top. induced in Trypanosoma cruzi metacyclic trypomastigotes during the invasion
Microbiol. Immunol. 117, 1–22. of mammalian cells. Braz. J. Med. Biol. Res. 33, 269–278.
Woolsey, A.M., Burleigh, B.A., 2004. Host cell actin polymerization is Zacks, M.A., Garg, N., 2006. Recent developments in the molecular, biochemical and
required for cellular retention of Trypanosoma cruzi and early functional characterization of GPI8 and the GPI-anchoring mechanism. Mol.
association with endosomal/lysosomal compartments. Cell. Microbiol. 6, Membr. Biol. 23, 209–225.
829–838. Zago, M.P., Barrio, A.B., Cardozo, R.M., Duffy, T., Schijman, A.G., Basombrio, M.A.,
Woolsey, A.M., Sunwoo, L., Petersen, C.A., Brachmann, S.M., Cantley, L.C., Burleigh, 2008. Impairment of infectivity and immunoprotective effect of a LYT1 null
B.A., 2003. Novel PI 3-kinase-dependent mechanisms of trypanosome invasion mutant of Trypanosoma cruzi. Infect. Immun. 76, 443–451.
and vacuole maturation. J. Cell Sci. 116, 3611–3622. Zhang, L., Tarleton, R.L., 1999. Parasite persistence correlates with disease severity
Yoshida, N., 2006. Molecular basis of mammalian cell invasion by Trypanosoma and localization in chronic Chagas’ disease. J. Infect. Dis. 180, 480–486.
cruzi. An. Acad. Bras. Cienc. 78, 87–111. Zhong, L., Lu, H.G., Moreno, S.N.J., Docampo, R., 1998. Tyrosine phosphate hydrolysis
Yoshida, N., Cortez, M., 2008. Trypanosoma cruzi: parasite and host cell signaling of host proteins by Trypanosoma cruzi is linked to cell invasion. FEMS Microbiol.
during the invasion process. Subcell. Biochem. 47, 82–91. Lett. 161, 15–20.

You might also like