Download PDF
ads:
Vanessa Diniz Atayde
Estudo das proteínas da família gp82 das
formas metaclicas do Trypanosoma cruzi
Tese apresentada à Universidade
Federal de o Paulo, Escola Paulista
de Medicina, para obtenção do Título
de Doutor em Ciências.
São Paulo
2008
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
Vanessa Diniz Atayde
Estudo das proteínas da família gp82 das
formas metaclicas do Trypanosoma cruzi
Orientadora: Dra. Nobuko Yoshida
Tese apresentada à Universidade
Federal de o Paulo, Escola Paulista
de Medicina, para obtenção do Título
de Doutor em Ciências.
São Paulo
2008
ads:
Atayde, Vanessa Diniz
Estudo das proteínas da família gp82 das formas metacíclicas do
Trypanosoma cruzi. / Vanessa Diniz Atayde. - - São Paulo, 2008.
xii, 77p.
Tese (Doutorado) Universidade Federal de o Paulo. Escola
Paulista de Medicina. Programa de s-Graduação em Microbiologia,
Imunologia e Parasitologia.
Título em inglês: Study of gp82 family proteins of Trypanosoma
cruzi metacyclic forms.
1.Trypanosoma cruzi 2.Formas metacíclicas 3.Invasão celular
4.Clone CL-14 5.Proteínas da família gp82 6.Melanoma 7.Apoptose
i
Trabalho realizado na Disciplina de Parasitologia do Departamento
de Microbiologia, Imunologia e Parasitologia da Universidade
Federal de São Paulo, Escola Paulista de Medicina, com auxílios
financeiros concedidos pela Fundação de Amparo à Pesquisa do
Estado de São Paulo (FAPESP) e pelo Conselho Nacional de
Desenvolvimento Cienfico e Tecnológico (CNPq).
ii
À Dra. Nobuko Yoshida, que me ensinou algo essencial
entre os mil milhões” de coisas que me ensinou nesses
quase dez anos de convivência: a cientista que eu quero
me tornar um dia. (Sempre trabalhando intensamente!!!)
iii
Aos meus pais, Hely e Al, que SEMPRE me deram total
apoio e compreensão. À Dani, minha eterna irmãzinha.
NUNCA teria conseguido sem eles.
iv
Agradecimentos
Foram quase dez anos na Disciplina de Parasitologia da UNIFESP. Aqui eu
obtive conhecimento, formação e exemplos que seguirei por toda a vida. Aqui eu
descobri que o nada que eu realmente queira fazer além da ciência. Aqui eu fiz
amigos inesquecíveis que estarão sempre no meu coração.
Em primeiro lugar, agradeço à Dra. Nobuko Yoshida, por tudo que fez por mim
nesses anos de trabalho e aprendizado intensos. Tenho um enorme respeito e
admiração por ela, que é um dos meus maiores exemplos de vida. Sei que dificilmente
encontrarei alguém parecido daqui para frente, mas fico muito feliz por ter aprendido
tanto no tempo em que convivemos.
Agradeço a todos os Professores do Departamento, especialmente ao Dr.
Maurício Rodrigues e ao Dr. Renato Mortara, que foram muito importantes em
diferentes momentos da minha formação.
Agradeço à Ciça, minha melhor amiga, minha irmã, minha querida. Obrigada
pela paciência, pelas risadas, pelas longas conversas (e repetitivas...), e até mesmo
pelas broncas muitas vezes merecidas. Estou chegando amiga!
Agradeço à Fanny, minha grande (literalmente) e querida amiga. Suas histórias
malucas me fizeram rir até nos dias mais diceis. Uma das melhores pessoas que
conheci na minha vida. estou com saudades.
Agradeço aos amigos do laboratório que contribuíram muito, cada um a seu
modo: Daniele, Renata S., Renata B., Marjorie, bio, Cláudio, Emanuelle, Carol,
Miriam, Simone, Érico, Carlinhos, Carla, lvia, Meire, Dú, Rafa, Silene, Gisa, Denis,
Josie, Paulo, Sheila, Jú, Mário....Espero encontrar todos vocês novamente!
v
Agradeço aos novos amigos, mas não menos importantes, Andrea Midory,
Fernando Maeda e Dani Staquicini. Aprendi muito com esses japinhas que têm um
grande caráter e um enorme coração. Tenho certeza que nossa amizade é e sempre
se 100%!
Agradeço aos amigos chilenos: Patcio Porcile, Sérgio Málaga, Ivan Neira,
Mauro Cortez (o Javi..., meu amigão!), Esteban Cordero e Charles Covarrubias.
Sentirei saudades da nossa convivência (principalmente das nossas festas...). Tenho
que agradecer especialmente ao Patricio Manque, meu primeiro contato com a
pesquisa. Obrigada pela sua amizade e por sempre acreditar na minha capacidade.
Agradeço à Mércia e à Regiane, sempre prestativas, amigas e muito elegantes!
E aos funcionários da disciplina: Dona Joaninha, Kleber, Luis, Lima, Adilson, Cidinha,
Dona Fátima e Seu Adalton. Um agradecimento especial à Sandra, minha fofa!
Por fim, agradeço à minha família muito unida e feliz, que tornou possível a
realização dessa tese. Com o apoio deles, tudo acabou ficando mais fácil. Amo vocês!
vi
Navegadores antigos tinham uma frase gloriosa:Navegar
é preciso, viver não é preciso.
Quero para mim o espírito desta frase, transformada a
forma para a casar com o que eu sou: Viver o é necessário; o
que é necessário é criar.
Não conto gozar a minha vida; nem em gozá-la penso. Só
quero torná-la grande, ainda que para isso tenha de ser o meu
corpo e a minha alma a lenha desse fogo.
Só quero torná-la de toda a humanidade; ainda que para
isso tenha de a perder como minha.
Cada vez mais assim penso. Cada vez mais ponho na
essência anímica do meu sangue o propósito impessoal de
engrandecer a pátria e contribuir para a evolução da humanidade.
É a forma que em mim tomou o misticismo da nossa Raça.”
Fernando Pessoa (Obra Poética)
vii
Índice
Resumo ix
Abstract xi
Introdução: Trypanosoma cruzi e Doença de Chagas
1. Doença de Chagas 1
2. Trypanosoma cruzi, agente etiológico da Doença de Chagas 4
3. Cepas e grupos do Trypanosoma cruzi 7
4. Infecção oral pelo Trypanosoma cruzi 10
5. Mecanismos de invasão celular pelas formas metacíclicas do
Trypanosoma cruzi 12
5.1. Gp82: glicoproteína de 82kDa expressa na superfície das
formas metacíclicas envolvida na invasão celular 13
5.2. Outras moléculas de superfície das formas metacíclicas
envolvidas na invasão celular 16
5.3. Vias de sinalização ativadas durante a invasão das formas
metacíclicas do Trypanosoma cruzi 20
Capítulo 1 Bases moleculares da avirulência do clone CL-14 do
Trypanosoma cruzi
Artigo: Molecular basis of non-virulence of Trypanosoma cruzi
clone CL-14” 24
viii
Capítulo 2 Expressão e localização celular de moléculas da família
gp82 em formas metacíclicas do Trypanosoma cruzi
Artigo: Expression and cellular localization of molecules of the
gp82 family in Trypanosoma cruzi metacyclic trypomastigotes” 28
Capítulo 3 Indução de apoptose em lulas de melanoma pela gp82
em sua forma recombinante
Artigo: A recombinant protein based on Trypanosoma cruzi surface
molecule gp82 induces apoptotic cell death in melanoma cells” 39
Discussão 50
Referências Bibliográficas 60
ix
Resumo
Nosso principal objetivo foi caracterizar a família gp82 de protnas,
codificada pela família gênica gp82, que é parte da grande família multigênica
gp85/trans-sialidase do T. cruzi. Até o icio desse estudo, somente a gp82 de
superfície, envolvida na invasão celular das formas metacíclicas da cepa CL e
identificada pelo anticorpo monoclonal (MAb) 3F6, havia sido descrita.
Na primeira parte, investigamos as bases da avirulência in vivo e in vitro
das formas metacíclicas do clone CL-14 do T. cruzi, derivado da cepa CL.
Descobrimos que a baixa infectividade desses parasitas está associada com a
reduzida expressão da gp82 na superfície, reforçando o papel central da
molécula na infecção pelo T. cruzi. Além disso, os dados sugeriram que devido
à deficiência da gp82, as formas metacíclicas do clone CL-14, como as da cepa
G, utilizam preferencialmente as glicoproteínas gp35/50 para interagir com a
célula hospedeira.
Na segunda parte, analisamos a expressão e a localização de moléculas
da família gp82 em formas metacíclicas das cepas G e CL do T. cruzi. Para
isso, isolamos um novo membro (clone C03) que, em contraste com o clones
de cDNA representantes da família gp82 previamente descritos, J18 (cepa G) e
R31 (cepa CL), não contém o epítopo do MAb 3F6. Na análise comparativa da
seqüência de aminoácidos do clone C03, observamos 59,1% de identidade
com o clone J18 e 60,2% de identidade com o clone R31. Além disso, quando
alinhamos as seências de aminoácidos dos clones C03 e Tc-85 (gp85 das
formas tripomastigotas de cultura de tecido), observamos 57,2% de identidade,
indicando que esse clone também é parte da família gp85/trans-sialidase.
Utilizando os anticorpos policlonais anti-J18 e anti-C03, e o MAb 3F6,
mostramos que membros da família gp82 são expressos em formas
metacíclicas na superfície e intracelularmente, e que alguns desses membros
têm localização diferenciada em parasitas dos grupos T. cruzi I e II. Ao
contrário da gp82 reativa com o MAb 3F6, específica das formas metacíclicas,
x
outras moléculas da família foram detectadas também nas formas
tripomastigotas de cultura de tecido e amastigotas pelos anticorpos policlonais.
Na terceira parte, investigamos o efeito da gp82 em sua forma
recombinante (J18), sobre a linhagem de melanoma murino Tm5. Observamos
que J18 liga-se a essas células induzindo a despolimerização dos filamentos
de actina similar à citocalasina-D, droga indutora de apoptose em células de
mamíferos. Em vista disso, fizemos uma análise comparativa das alterações no
citoesqueleto de actina e eventos característicos de morte celular por apoptose
nas células Tm5 tratadas com J18, e na linhagem de melanócitos melan-a, da
qual Tm5 é derivada. Mostramos que J18 tem função indutora de apoptose
somente sobre as células do melanoma Tm5. Descrevemos assim uma
propriedade de um membro da família gp82, nunca antes descrita para nenhum
outro componente molecularmente definido do T. cruzi.
xi
Abstract
The main goal of this work was to characterize the proteins encoded by
the gp82 gene family, which is part of the large T. cruzi gp85/trans-sialidase
multigenic family. Previous studies had focused on the surface gp82, which is
engaged in the cell invasion by CL strain metacyclic forms and is identified by
monoclonal antibody (MAb) 3F6.
Firstly, we investigated the molecular basis of the non-virulence of T.
cruzi clone CL-14 metacyclic forms in vivo and in vitro. We found that the low
infectivity of these parasites is associated with reduced expression of surface
gp82, reinforcing the role of this molecule in T. cruzi infection. In addition, our
data suggested that instead of gp82, metacyclic forms of clone CL-14, like G
strain, preferentially engage gp35/50 glycoproteins to interact with the host cell.
Secondly, we analyzed the expression and cellular localization of
molecules of the gp82 family in T. cruzi metacyclic forms of G and CL strains.
We cloned a new member of this family, designated C03, which lacks the MAb
3F6 epitope, in contrast to the previously described gp82 cDNA clones J18 (G
strain) and R31 (CL strain). The predicted amino acid sequence of the C03
clone displayed 59,1% identity to J18 clone and 60,2% to R31 clone. When the
amino acid sequences of C03 and Tc-85 (tissue culture trypomastigote gp85)
clones were aligned, the identity was 57,2%, indicating that this new clone
belongs to the gp85/trans-sialidase family. Using anti-J18 and anti-C03
polyclonal antibodies, as well as MAb 3F6, we demonstrated that members of
the gp82 family are localized on the cell surface and intracellularly in metacyclic
forms, and some members have different cellular localization in parasites from
T. cruzi I and II groups. As opposed to metacyclic stage-specific gp82 identified
by MAb 3F6, other gp82 family molecules were also detected in tissue culture
trypomastigotes and amastigotes by polyclonal antibodies.
Thirdly, we investigated the effect of gp82, as a recombinant protein
(J18), on the murine melanoma lineage Tm5. We observed that J18 binds to
xii
these cells disrupting actin filaments similarly to cytochalasin-D, drug that
induces apoptosis in mammalian cells. Based on these information, we
comparatively analyzed alterations in actin cytoskeleton and apoptotic events in
J18-treated Tm5 cells, and in the melanocyte lineage melan-a, from which Tm5
is derived. J18 selectively induced apoptosis in Tm5 melanoma cells. Our
results show an activity of a protein from gp82 family that has not been
described for any other T. cruzi molecule.
1
Introdução: Trypanosoma cruzi e Doença de Chagas
1. Doença de Chagas
A doença de Chagas ou tripanossomíase americana foi descoberta pelo
brasileiro Carlos Chagas (Fig.1) em 1909, às margens do Rio São Francisco
(MG), enquanto combatia uma epidemia de malária que atrasava a construção
da ferrovia Central do Brasil. Com a identificação do agente etiológico e a
descrição da moléstia, Chagas foi o primeiro cientista a apontar a importância
sócio-econômica da doença no país. Além disso, fundou as bases da pesquisa
básica realizada em todo o século XX até os dias atuais, essencial no
entendimento das interações parasita-hospedeiro.
Figura 1. Carlos Chagas (FIOCruz).
Eram os “anos de ouro” da medicina tropical, ciência prestigiada na
época em razão da necessidade de manter íntegra a saúde dos soldados
brancos e colonizadores do final do século XIX, assolados pelas doenças
tropicais. O Império Britânico, que se estendia a diversos países da África,
transformava a medicina tropical em negócio comercial e militar (Worboys,
1993). Nesse período emergiram os caçadores de parasitas”, como Ross e
2
Laveran, que ganharam o prêmio Nobel em 1902 e 1907 por suas descobertas
em malária. Chagas foi nomeado em 1913 e 1921 sem ser premiado, pois a
comunidade científica não se convencia da importância e do impacto da sua
descoberta (Coutinho et al., 1999).
Atualmente, a doença de Chagas é considerada uma das doenças
tropicais negligenciadas causadas por protozrios, junto com a
tripanossomíase africana e a leishmaniose (Boutayeb et al., 2007). Segundo a
Organização Mundial de Saúde (WHO, 2008), as doenças tropicais
negligenciadas afetam mais de 1 bilhão de indiduos em todo o mundo,
causando aproximadamente 500.000 mortes anuais. A maior parte dos
doentes, que vivem em situações de extrema pobreza, encontra-se co-
infectada com mais de um patógeno, já que na maioria dos países afetados
ocorre concomitantemente mais de uma doença (Reddy et al., 2007) (Fig.2). As
interações entre os patógenos que co-infectam um indivíduo podem ser
sinérgicas ou competitivas e, sendo assim, uma doença pode interferir no
impacto geral da outra (Buck et al., 1978, Petney & Andrews, 1998).
Figura 2. Ocorrência global das doeas tropicais negligenciadas. Cinza: uma
doença, azul: duas, verde: três, amarelo: quatro, laranja: cinco, vermelho: seis
(gnntdc.sabin.org).
3
Apesar da grande importância sócio-econômica, poucos esforços são
despendidos no combate às doenças tropicais negligenciadas, provavelmente
porque não afetam países ricos. Nos últimos 30 anos, somente 1% dos
investimentos mundiais no desenvolvimento de novas drogas foram
direcionados ao tratamento dessas doenças (Beyrer et al., 2007, Reddy et al.,
2007).
A doença de Chagas é um grave problema de saúde pública de pses
da América Latina, sendo enmica do norte do xico ao sul da Argentina
(Fig.3). Segundo a WHO (2005), existem aproximadamente 13 milhões de
indiduos infectados e cerca de 14.000 mortes anuais, causadas
principalmente pelos danos irreversíveis ao coração, característicos da doença.
Figura 3. Endemia global da doea de Chagas (Morel & Lazdins, 2003).
Populações rurais vivendo em habitações de baixa qualidade, onde o
inseto vetor se estabelece, são os principais alvos da doença de Chagas. Em
países do cone sul da América Latina como Argentina, Brasil e Chile, medidas
de controle da proliferação do inseto vetor têm sido eficientes na diminuição da
incidência anual, que passou de 700.000-800.000, estimativa de 1990, para
4
200.000 casos (WHO, 2005). Além disso, dos 100 milhões de indiduos
vivendo sob risco de infecção em 1990, hoje existem apenas 40-75 milhões
(Schofield et al., 2006, Coura, 2007). Entretanto, essas medidas de controle
não são eficazes contra os insetos silvestres ou peridomiciliares, que podem
eventualmente reinfestar as casas ou contaminar os alimentos (Schofield et al.,
2006).
2. Trypanosoma cruzi, agente etiológico da Doença de Chagas
O Trypanosoma cruzi é um protozoário flagelado pertencente à ordem
Kinetoplastidae, que se caracteriza pela presença de cinetoplasto, única
mitocôndria que contém uma complexa rede de DNA em seu interior. É
digenético, pois necessita de um hospedeiro vertebrado e de um hospedeiro
invertebrado para completar seu ciclo de vida. Os hospedeiros invertebrados
são os insetos triatomíneos da família Reduviidae, conhecidos popularmente
como “barbeiros”. Os membros dessa família que têm maior imporncia
epidemiológica são o Triatoma infestans, Triatoma brasiliensis, Triatoma
dimidiata, Rhodnius prolixus e Panstrongylus megistus (Schofield, 2000).
Três formas do T. cruzi são identificadas durante seu ciclo de vida:
amastigotas e epimastigotas, replicativas, e tripomastigotas (sanguíneas ou
metacíclicas), infectivas (Fig.4).
Figura 4. Formas tripomastigota, amastigota e epimastigota (uta.edu).
5
A transmissão vetorial inicia-se quando o inseto vetor, após o repasto
sanguíneo, defeca liberando as formas tripomastigotas metacíclicas. Essas
formas penetram o hospedeiro vertebrado pela lesão da picada ou diretamente
pelas mucosas, onde invadem as células das camadas subjacentes. Dentro
das células, as formas metacíclicas encontram-se em um compartimento
derivado da membrana plasmática e de lisossomos, o vacúolo parasitóforo, de
onde escapam e diferenciam-se em amastigotas. As formas amastigotas
replicam-se no citoplasma e em seguida diferenciam-se em tripomastigotas,
liberados após a ruptura celular. Os tripomastigotas infectam células vizinhas
ou caem na corrente sanguínea, espalhando a infecção aos diferentes órgãos e
tecidos. As formas tripomastigotas sanguíneas circulantes são ingeridas pelo
barbeiro durante o repasto e diferenciam-se em formas epimastigotas, que se
replicam na porção média do seu intestino e migram à porção final, onde se
diferenciam em formas tripomastigotas metacíclicas, completando o ciclo
(Fig.5).
Figura 5. Ciclo de vida do T. cruzi (WHO).
6
Há três ciclos de transmissão vetorial do T. cruzi. O ciclo de maior
importância epidemiológica é o doméstico, já que perpetua a infecção nos
seres humanos. No ciclo silvestre, participam triatomíneos que, uma vez
contaminados, infectam roedores, marsupiais e outros animais silvestres. O
terceiro ciclo é o peridoméstico, do qual participam roedores, marsupiais, gatos
e cães que entram e saem das residências, e os insetos silvestres atraídos às
casas pela disponibilidade de alimento. Esse ciclo serve de ligação entre os
ciclos doméstico e silvestre (Rey, 2002). Um estudo recente mostrou que em
91,5% das vilas rurais da região oeste da Venezuela, onde a doença de
Chagas é enmica, foram encontrados cães infectados com T. cruzi,
convivendo diretamente com humanos infectados em casas que continham
ninhos do inseto vetor. Isso mostra que animais domésticos são importantes
fatores de risco na transmissão da doença de Chagas aos humanos (Crisante
et al., 2006).
Além da transmissão vetorial, também ocorrem as transmissões
connita, transfusional ou por transplante de órgãos. Esses tipos de
transmissão têm levado a doença de Chagas às regiões não-enmicas, como
a América do Norte, através da migração de indiduos infectados (Buekens et
al., 2008). A transmissão oral, que será discutida adiante, é de grande
importância na constante manutenção do ciclo silvestre e tornou-se evidente
devido aos casos de infecção humana por ingestão de alimentos contaminados
com o parasita.
A doença é caracterizada por uma fase aguda, com sintomas que
surgem alguns dias após a contaminação, e por uma fase crônica, com
sintomas que surgem após um período latente de 10 a 30 anos. A lesão inicial,
ou chagoma de inoculação, chama particularmente a atenção quando ocorre
no olho (sinal de Romaña) (Fig.6). Os sintomas da fase aguda são oriundos da
ruptura das células infectadas e da presença de parasitas no sangue, o que
desencadeia uma resposta inflamatória com febre, edema, linfadenopatia e
miocardite (Rey, 2002). Em 95% dos casos, essa fase é assintomática e
somente 5% dos indiduos sintomáticos morrem. As lesões da fase crônica
afetam irreversivelmente órgãos como o coração, esôfago e intestino, pois uma
7
fibrose difusa ocupa o lugar das áreas inflamadas. De 30% dos pacientes
crônicos que desenvolvem manifestações clínicas, 95% têm o coração afetado.
Infecções pelo HIV podem reativar a doença em pacientes crônicos,
evidenciando a persistência do parasita no hospedeiro durante essa fase
(Punnukollu et al., 2007, Teixeira et al., 2008). No entanto, a maioria dos
infectados tem a forma indeterminada da doença de Chagas, sem nunca
desenvolver a doença crônica (Rey, 2002).
Figura 6. Sinal de Romaña (tmcr.usuhs.mil).
O tratamento, quando o doente tem acesso a ele, é insatisfatório devido
à sua ineficácia no estágio crônico, além dos diversos efeitos tóxicos. As
drogas utilizadas são o nifurtimox e o beznidazol, que têm efeitos significativos
na fase aguda da doença, com mais de 80% de cura parasitológica em
pacientes tratados (Rassi et al., 2000). Porém, sua eficácia varia em diferentes
regiões geográficas, provavelmente devido às variações de sensibilidade das
cepas do parasita aos medicamentos, conforme demonstrado
experimentalmente e no homem (Brener et al., 1976, Andrade et al., 1992).
3. Cepas e grupos do Trypanosoma cruzi
Os isolados ou cepas do T. cruzi, provenientes de vários hospedeiros
selvagens ou domésticos, vertebrados ou invertebrados, apresentam-se
8
altamente heterogêneos segundo critérios morfológicos, genéticos, bioquímicos
e clínicos (Devera et al., 2003). O elevado grau de diversidade do parasita
pode estar associado à sua necessidade de adaptação e sobrevivência em
diferentes hospedeiros.
Analisando-se padrões de isoenzimas, as cepas do T. cruzi foram
inicialmente divididas em três zimodemas, Z1, Z2 e Z3 (Miles et al., 1980). Em
seguida, Dvorak et al. (1982) mostraram que as cepas possuem diferentes
quantidades de DNA total por parasita, sugerindo que a heterogeneidade
biológica poderia ser resultante de diferenças genéticas. A partir desse
momento, o maior desafio foi identificar marcadores moleculares que se
correlacionassem com o comportamento variável das cepas estudadas.
Em 1996, Souto et al. agruparam os isolados do T. cruzi em duas
linhagens, através da análise da seência do rDNA 24S e das seqüências
internicas dos genes de mini-éxons. Amostras de diferentes hospedeiros de
vários países da América Latina foram classificadas em linhagem 1 (Z2) ou
linhagem 2 (Z1). Cepas pertencentes ao Z3 não puderam ser classificadas,
uma vez que apresentaram seqüências incomuns (Mendonça et al., 2002). Em
1999, num Simpósio Internacional, a comunidade científica sugeriu a
padronização da nomenclatura: o que foi originalmente classificado como
linhagem 1 e linhagem 2, foi renomeado como grupo T. cruzi II e grupo T. cruzi
I, respectivamente (Satellite-Meeting, 1999). Posteriormente, o grupo T. cruzi II
foi subdividido em cinco subgrupos (IIa-IIe), resultantes de fusões genéticas
com sinais de recombinação e heterozigose (Brisse et al., 2000, Westenberger
et al., 2005). Em outras análises, o Z3 foi caracterizado como parte de um
grupo ancestral chamado T. cruzi III (Pedroso et al., 2007). O clone CL-Brener
do projeto genoma (El-Sayed et al., 2005), que contém marcadores de DNA
dos grupos T. cruzi I, II e III sendo considerado um clone híbrido, foi agrupado
ao sub-grupo IIe (Elias et al., 2005, Pedroso et al., 2007, Cerqueira et al.,
2008).
Estabelecidos os critérios de classificação dos isolados, iniciaram-se os
estudos que tinham como objetivo associar os grupos genéticos com
9
características epidemiológicas e clínicas da doença de Chagas. Em muitos
estudos, a variabilidade nos sintomas foi atribda à diversidade das cepas
(Baptista et al., 2006). Uma explicação para essa diversidade seria a expressão
diferencial de genes específicos. No caso das formas metacíclicas, diferentes
cepas apresentam repertórios de superfície distintos e interagem com as
células-alvo de modos particulares (Yoshida, 2006).
Em análise epidemiológica de cepas de animais silvestres e domésticos,
vetores e humanos de áreas endêmicas e não-endêmicas do Brasil, Colômbia
e Bolívia, observou-se a associação do grupo T. cruzi II com o ciclo doméstico
e do grupo T. cruzi I, composto principalmente por isolados de gambás, com o
ciclo silvestre. Esses grupos foram considerados filogeneticamente divergentes
(Briones et al., 1999, Zingales et al., 1999). Cepas pertencentes ao grupo T.
cruzi III, freqüentemente encontradas em triatomíneos silvestres, representam
um reservatório constitutivo de infecção humana na Amazônia (Pedroso et al.,
2007).
O grupo T. cruzi II possui alta incincia em doentes crônicos de regiões
enmicas do estado de Minas Gerais (Lages-Silva et al., 2006). Esse grupo
também é freqüentemente identificado em pacientes crônicos da Argentina e
Chile, podendo ser o único que causa a doença nesses países (Di Noia et al.,
2002). o grupo T. cruzi I é endêmico no norte da América do Sul, incluindo a
Amazônia, onde a doença de Chagas é emergente devido às microepidemias
por infecção oral e aos doentes com sintomas severos (Coura et al., 2002,
Miles et al., 2003, Teixeira et al., 2006). Recentemente, uma cepa do grupo T.
cruzi I (cepa José-IMT) foi isolada de um paciente crônico com graves sintomas
cardíacos (Teixeira et al., 2006), mostrando que esse grupo também pode
causar a doença crônica. Esse paciente provinha de uma vila rural do estado
da Paraíba, próxima a Catolé do Rocha onde, em 1986, 26 pessoas se
contaminaram num surto de infecção oral após a ingestão de caldo de cana,
com um caso fatal (Shikanai-Yasuda et al., 1991).
10
4. Infecção oral pelo Trypanosoma cruzi
Análises filogenéticas sugerem que a infecção pelo T. cruzi iniciou-se
milhões de anos em animais silvestres, da maneira que ainda persiste em
regiões de florestas selvagens como a Amazônia (Briones et al., 1999, Coura,
2007). Apesar de existirem relatos de infecção humana em múmias de até
9.000 anos (Aufderheide et al., 2004), a doença endêmica somente se
estabeleceu com o desmatamento das florestas, nos últimos 300 anos (Aragão,
1983). Com a colonização de áreas próximas ao homem pelos triatomíneos,
que se adaptaram ao novo nicho alimentando-se do sangue humano e de
animais domésticos, a infecção tornou-se uma zoonose, a doença de Chagas
(Coura, 2007).
No ciclo silvestre, a contaminação oral com o T. cruzi manm constante
a infecção dos animais, devido ao hábito de se alimentarem de pequenos
mamíferos e insetos (Jansen et al., 1997). Os gambás (Fig.7) têm um
importante papel nesse ciclo, já que não somente carregam as formas
sanguíneas, mas também são reservatórios de formas metacíclicas,
encontradas no lúmen de sua glândula anal (Deane, 1984). Por outro lado,
animais domésticos podem invadir o ambiente selvagem para caçar,
infectarem-se por via oral e levarem a infecção para o ambiente domiciliar
(Coura, 2007).
Figura 7. Gams (Didelphis marsupialis).
11
No caso do homem, com o controle da proliferação do inseto vetor e dos
bancos de sangue, a infecção oral é hoje a mais importante forma de
transmissão da doença de Chagas (Coura, 2006). Mais da metade dos casos
de doença aguda, registrados entre 1968 e 2000 na Amazônia brasileira, são
atribuídos a microepidemias causadas por infecção oral após a ingestão de
alimentos contaminados com o T. cruzi (Coura et al., 2002). Esse tipo de
infecção está em destaque desde 2005, quando foram relatados 26 casos no
estado de Santa Catarina, resultantes da ingestão de caldo de cana com o
parasita. Os focos da contaminação foram triatomíneos moídos ou somente
suas fezes contendo formas metacíclicas, especializadas na invasão do epitélio
da mucosa gástrica em camundongos (Hoft, 1996, Hoft et al., 1996). Outros
eventos semelhantes foram relatados nos estados do Amazonas, Amapá e
Pará (Coura et al., 2002, Promed-mail, 2007) (Fig.8). Os alimentos ingeridos
pelos seres humanos tamm podem ser contaminados pela secreção da
glândula anal de gambás. Em 1986, na Paraíba (Fig.8), onde diversas pessoas
adquiriram a doença aguda após ingerir caldo de cana, gambás altamente
infectados foram encontrados (Shikanai-Yasuda et al., 1991).
Figura 8. Microepidemias de doea de Chagas por infecção oral no Brasil.
12
5. Mecanismos de invasão celular pelas formas metacíclicas do
Trypanosoma cruzi
O T. cruzi invade diversos tipos celulares, incluindo células epiteliais,
musculares, macrófagos e fibroblastos, e esse é um fator essencial no
estabelecimento da infecção do hospedeiro vertebrado. Em ensaios in vitro,
inúmeros isolados do parasita e linhagens de lulas de mamíferos, m sido
utilizados (Fig.9). Esses ensaios são imprescindíveis no estudo do processo de
invasão, onde interagem vários componentes do parasita e da lula
hospedeira, ativando vias de sinalização bidirecionais que promovem a
internalização (Burleigh & Andrews, 1998, Mortara et al., 2005, Yoshida, 2006).
É importante lembrar que as estratégias de invasão celular utilizadas pela
forma tripomastigota de cultura de tecido (referente à forma sangnea) e pela
forma metacíclica são distintas. Apesar de ambas serem infectivas, cada uma
possui seu repertório de moléculas de superfície e interage de maneira
particular com as lulas-alvo.
Figura 9. Cepas ou clones do T. cruzi e células de mamíferos utilizados em
ensaios de invao (adaptado de Yoshida, 2006).
13
Nas infecções naturais, as formas metacíclicas são as primeiras a
entrarem em contato com as células do hospedeiro. A capacidade infectiva
dessas formas difere entre as cepas, devido à expressão diferencial de
moléculas de superfície. Normalmente, cepas pertencentes ao grupo T. cruzi II
são mais infectivas in vitro do que cepas pertencentes ao grupo T. cruzi I (Ruiz
et al., 1998, Neira et al., 2002, Yoshida, 2006). In vivo, essa correlação nem
sempre existe, que fatores do hospedeiro devem ser considerados
(Covarrubias et al., 2007).
Em estudos com formas metacíclicas de cepas muito infectivas, como
por exemplo a cepa CL (Fig.10), têm-se confirmado o papel central da molécula
de superfície gp82 na invasão celular, in vitro e in vivo (Ramirez et al., 1993,
Santori et al., 1996, Ruiz et al., 1998, Yoshida et al., 2000, Neira et al., 2003).
Figura 10. Forma metaclica da cepa CL marcada com anticorpo anti-gp82.
5.1. Gp82: glicoproteína de 82kDa expressa na superfície das formas
metacíclicas envolvida na invasão celular
Responvel pela interação inicial das formas metacíclicas com a lula
hospedeira, a gp82 promove a invasão de cepas altamente infectivas como a
cepa CL (Ramirez et al., 1993, Ruiz et al., 1998). Essa glicoproteína contém
oligossacarídeos N-ligados (Ramirez et al., 1993) e está inserida na membrana
14
plasmática via âncora de GPI, forma de ancoramento de muitas proteínas
eucarióticas (Schenkman et al., 1988, Cardoso de Almeida & Heise, 1993). A
porção glidica da gp82 não está envolvida na invasão, e o sítio de adesão
celular encontra-se no domínio central da molécula e é formado pela
justaposição de duas seqüências peptídicas carregadas separadas por outra
hidrofóbica (Santori et al., 1996, Manque et al., 2000).
As evidências da importância da gp82 no processo de invasão são: i) o
anticorpo monoclonal (MAb) 3F6, contra uma porção do domínio central da
gp82 adjacente ao sítio de adesão celular, inibe a entrada do parasita em
células não-fagocíticas (Ramirez et al., 1993); ii) tanto a gp82 nativa como a
sua forma recombinante inibem a penetração do parasita nalula hospedeira
(Ramirez et al., 1993, Santori et al., 1996); iii) a ligação da gp82 ao seu
receptor, ainda não identificado, leva à mobilização de Ca
+2
intracelular em
ambas as células (Dorta et al, 1995, Ruiz et al., 1998), evento essencial para a
invasão (Yakubu et al., 1994, Moreno et al., 1994, Tardieux et al., 1994).
Formas epimastigotas não induzem sinal intracelular de Ca
+2
(Tardieux et al.,
1994, Dorta et al., 1995), ao menos que estejam transfectadas com um vetor
de expressão carregando a seqüência de cDNA da gp82 (Manque et al., 2003).
Na infecção in vivo, a gp82 tem função central na invasão do epitélio da
mucosa gástrica, quando formas metacíclicas o inoculadas oralmente em
camundongos (Hoft, 1996, Hoft et al., 1996, Neira et al., 2003). A primeira
evidência do envolvimento da gp82 na infecção oral foi a inibição da
infectividade da cepa CL pelo tratamento com o MAb 3F6. Anticorpos contra
outras proteínas de superfície do T. cruzi não apresentaram o mesmo efeito.
Além disso, foi visto que a gp82 é capaz de ligar-se diretamente à mucina
gástrica, sendo possivelmente essa a primeira etapa no acesso das formas
metacíclicas às células da mucosa (Neira et al., 2003). No epitélio gástrico de
camundongos infectados oralmente, ninhos de amastigotas podem ser
observados (Fig.11). Uma particularidade relevante para a infecção oral é a
resistência da gp82 às condições do meio gástrico. Parasitas da cepa CL,
expostos ao suco gástrico in vivo ou tratados com pepsina in vitro se manm
15
infectivos às células de mamíferos, pois esses tratamentos não afetam a
integridade da gp82 (Neira et al., 2003, Cortez et al., 2006a).
As formas metacíclicas do T. cruzi também infectam a mucosa ocular.
Cães inoculados com essas formas desenvolvem altas parasitemias em
comparação com os inoculados com as formas sangneas, que causam
parasitemias muito baixas ou indetecveis (Bahia et al., 2002). Se há função
da gp82 na invasão da mucosa ocular ainda precisa ser determinado.
Figura 11. Ninhos de amastigotas da cepa CL no epitélio da mucosa gástrica de
camundongo infectado por via oral. Corte histológico corado com hematoxilina-
eosina (Yoshida, 2008).
Posteriormente, o papel da gp82 no estabelecimento da infecção oral foi
confirmado através do estudo dos isolados gp82-deficientes, 569 e 588 (Cortez
et al., 2003). Esses isolados produzem baixas parasitemias em camundongos
quando comparados com a cepa CL. Porém, in vitro, apresentam
comportamento similar à cepa CL, pois expressam a gp30. Essa molécula de
superfície é também reconhecida pelo MAb 3F6 e tem atividade sinalizadora de
Ca
+2
, mas não se liga à mucina gástrica eficientemente. Fica claro que a
invasão das células-alvo pelas formas metacíclicas pode ser promovida tanto
pela gp82 como pela gp30, que preserva o domínio de adesão celular da gp82.
16
Entretanto, a eficiente infecção da mucosa gástrica depende da gp82, devido à
sua capacidade de interagir com a mucina (Fig.12).
Figura 12. Interações entre as formas metaclicas e a mucosa strica durante
infecção oral (adaptado de Yoshida, 2008).
5.2. Outras moléculas de superfície das formas metacíclicas envolvidas
na invasão celular
As moléculas gp35/50 e gp90 também estão envolvidas na invasão das
formas metacíclicas. Como a gp82, essas moléculas ligam-se à lula
hospedeira de maneira receptor-dependente (Ruiz et al., 1998), porém até o
momento, nenhum dos receptores foi descrito. Cepas pouco infectivas, como a
cepa G, parecem utilizar preferencialmente as gp35/50 para interagir com as
células (Yoshida et al., 1989). Essas moléculas não são efetivas como a gp82,
provavelmente devido à sua reduzida atividade sinalizadora de Ca
+2
(Dorta et
17
al., 1995). Isolados pouco infectivos expressam também a gp90, que atua
como modulador negativo da invasão (Málaga & Yoshida, 2001).
Expressas na superfície de formas metacíclicas e de epimastigotas de
todos os isolados de T. cruzi analisados até o momento, as glicoproteínas
gp35/50 são moléculas do tipo mucina (Yoshida, 2006). O envolvimento dessas
moléculas na invasão celular foi determinado através da inibição da
infectividade de formas metacíclicas pelo MAb 10D8, direcionado a um epítopo
de carboidrato das gp35/50, ou pelas proteínas nativas purificadas (Yoshida et
al., 1989, Ruiz et al., 1993).
As glicoproteínas do tipo mucina são codificadas por uma família
multinica que representa aproximadamente 1% do genoma do T. cruzi (Di
Noia et al., 1998, Buscaglia et al., 2006). São ancoradas à membrana via GPI e
contém em sua estrutura úcares O-ligados a resíduos de treonina através de
uma N-acetilglicosamina, ao invés de uma N-acetilgalactosamina, como
normalmente é observado nas mucinas de vertebrados (Schenkman et al.,
1993, Acosta-Serrano et al., 2001). As moléculas de formas epimastigotas e
metacíclicas diferem entre si somente na porção lipídica da âncora GPI, tendo
a mesma porção glicídica (Acosta-Serrano et al., 1995). Dependendo da cepa
do T. cruzi, as mucinas contém resíduos de galactofuranose além dos resíduos
de galactopiranose (Previato et al., 1994, Acosta-Serrano et al., 1995, Acosta-
Serrano et al., 2001), apresentando um grande polimorfismo (Mortara et al.,
1992). O MAb 10D8, que reconhece as gp35/50 de isolados pouco invasivos
como as cepas G e Tulahuén, reage com etopos de galactofuranose,
enquanto o MAb 2B10 reage com epítopos de galactopiranose, presentes em
todas as cepas (Yoshida, 2006).
Na transferência de moléculas de ácido siálico do hospedeiro pela
enzima trans-sialidase, as gp35/50 atuam como principais aceptores. Essa
enzima, aproximadamente 30 vezes mais ativa em tripomastigotas de cultura
de tecido (TCT) do que em formas metacíclicas, transfere resíduos de ácido
siálico às galactoses disponíveis, e age como uma sialidase (neuraminidase)
na auncia desses açúcares aceptores (Schenkman et al., 1991a, Schenkman
18
& Eichinger, 1993, Acosta-Serrano et al., 2001). Dependendo da cepa do T.
cruzi, os resíduos de ácido siálico presentes nas gp35/50 atrapalham a
invasão. Por exemplo, o tratamento da cepa G com neuraminidase bacteriana,
aumenta a infectividade desta cepa. Além disso, as gp35/50 dessialiladas têm
maior atividade sinalizadora de Ca
+2
na lula hospedeira (Yoshida et al.,
1997). Em contraste, a infectividade da cepa CL não é afetada pelo tratamento
com neuraminidase (Yoshida et al., 1997), já que essa cepa utiliza a gp82 para
a sua internalização. Quanto a papel das moléculas gp35/50 na infecção oral
pelas formas metacíclicas, que o altamente resistentes à protlise, elas
devem proteger os parasitas da digestão pelo suco gástrico (Mortara et al.,
1992).
Ao contrário da gp82 e das mucinas gp35/50, a glicoproteína gp90,
específica de formas metacíclicas, atua como regulador negativo da invasão
celular. A função inibitória da gp90 sobre a invasão foi demonstrada em
experimentos onde oligonucleotídeos anti-senso, complementares à porção C-
terminal da molécula onde aparentemente está o sítio de adesão celular, foram
utilizados no tratamento de formas metacíclicas da cepa G. O tratamento
reduziu a expressão da molécula e aumentou significativamente a infectividade
dos parasitas (Málaga e Yoshida, 2001). A gp90, presente em múltiplas pias
no genoma do T. cruzi (Franco et al., 1993), contém oligossacarídeos N-ligados
e é ancorada à membrana via GPI (Schenkman et al., 1988, Yoshida et al.,
1990). Apesar de ligar-se à célula-alvo de maneira receptor-dependente, não
induz sinal intracelular de Ca
+2
(Ruiz et al., 1998). A infectividade das diferentes
cepas tem correlação inversa com a expressão de gp90, como por exemplo a
cepa CL, que expressa baixos níveis da molécula (Ruiz et al., 1998). Estudos
recentes com dois isolados de pacientes, 573 e 587, poderiam explicar a
severidade de alguns casos de doença aguda após a ingestão de formas
metacíclicas. Os dois isolados são pouco infectivos in vitro, porém, in vivo, a
cepa 573 produz altas parasitemias, semelhante à cepa CL. Isso se dá pelo
fato da gp90 da cepa 573 ser suscetível à degradação pelo suco gástrico
(Cortez et al, 2006a).
19
A análise da expressão de moléculas de superfície pelas formas
metacíclicas tem se revelado um bom todo para diferenciar cepas
pertencentes a grupos filogeneticamente divergentes, que diferem na
infectividade in vitro. Em análise de 10 cepas, de diferentes fontes e regiões
geográficas, observou-se que os parasitas altamente invasivos o deficientes
em gp90 e gp35/50 identificadas respectivamente pelos MAbs 1G7 e 10D8,
embora expressem formas variantes dessas moléculas, reconhecidas pelos
MAbs 5E7 e 2B10. Cepas pouco invasivas expressam gp90 reativa com os
MAbs 1G7 e 5E7, e gp35/50 reconhecidas pelos MAbs 2B10 e 10D8. A
expressão de gp82 reativa com o MAb 3F6 é ubíqua entre as cepas (Yoshida,
2006) (Fig.13).
Figura 13. Invao celular e expressão na superfície de cepas do T. cruzi
(adaptado de Yoshida, 2006).
20
5.3. Vias de sinalização ativadas durante a invasão das formas
metacíclicas do Trypanosoma cruzi
A eficiência na invasão depende da molécula de superfície que é
utilizada pelo parasita para interagir com a lula hospedeira. Após essa
interação, distintas vias de sinalização são ativadas em ambas as células,
como observado em estudos com as cepas G e CL, pertencentes a grupos
genéticos divergentes (Yoshida, 2006, Yoshida e Cortez, 2008). Essas vias
foram identificadas através do tratamento dos parasitas ou das células-alvo
com drogas já descritas para células de mamíferos (Neira et al., 2002, Ferreira
et al., 2006).
Formas metacíclicas da cepa G utilizam preferencialmente as moléculas
gp35/50 para interagir com a célula hospedeira (Yoshida et al., 1989, Yoshida
et al., 1997). Após a ligação da gp35/50 ao seu receptor, é ativada no parasita
a via de sinalização que envolve a adenilato ciclase, gerando AMP cíclico
(cAMP) e liberação de Ca
+2
de ácidocalciossomos (Neira et al., 2002), vacúolos
ácidos contendo um sistema de troca Ca
+2
/H
+
(Docampo et al., 1995) (Fig.14).
Por outro lado, a invasão celular pelas formas metacíclicas da cepa CL é
mediada pela gp82 (Ramirez et al., 1993). Após a ligação da gp82 ao seu
receptor, é ativada no parasita a via de sinalização que envolve a fosfolipase C
(PLC), com geração de trifosfato de inositol (IP
3
), e liberação de Ca
+2
de
compartimentos IP
3
-dependentes (Yoshida et al., 2000, Neira et al., 2002). Em
amastigotas, epimastigotas ou TCT, a liberação de Ca
+2
dependente de IP
3
não
é detectada (Moreno et al., 1994, Docampo et al., 1995). Além disso, após a
interação da gp82 com alula hospedeira, é fosforilada no parasita uma
proteína de 175 kDa, de função desconhecida (Favoreto et al., 1998, Neira et
al., 2002) (Fig.14). Os tripanossomatídeos Leishmania, Trypanosoma brucei e
Trypanosoma cruzi possuem um grande repertório de protnas quinases,
compreendendo aproximadamente 2% de seus genomas, sugerindo que a
fosforilação tem função chave na biologia desses parasitas. Entretanto, não
21
possuem proteínas tirosino-quinases típicas de eucariotos (El-Sayed et al.,
2005, Parsons et al., 2005).
Figura 14. Vias de sinalização ativadas nas formas metacíclicas das cepas G e
CL durante a invasão.
Na célula-hospedeira, duas vias de sinalização distintas são ativadas
durante a invasão de formas metacíclicas das cepas G ou CL (Cortez et al.,
2006b, Ferreira et al., 2006). A via ativada pela cepa G envolve o cAMP,
enquanto que, na via ativada pela cepa CL, estão envolvidas a PI3-quinase
(PI3-K) e a proteína quinase C (PKC), com liberação de Ca
+2
intracelular IP
3
-
dependente (Ferreira et al., 2006). Muitos componentes das vias ativadas,
tanto no parasita como na célula hospedeira, ainda precisam ser estudados.
22
Muitos patógenos intracelulares subvertem vias de sinalização da lula
hospedeira em seu próprio benefício, como por exemplo, as bactérias Shigella
e Escherichia coli enteroinvasiva, que induzem o remodelamento do
citoesqueleto celular durante a invasão (Bourdet-Sicard et al., 2000, Cossart &
Sansonetti, 2004). Ferreira et al. (2006) observaram que a invasão de formas
metacíclicas das cepas G ou CL é dependente ou independente do
citoesqueleto de actina da lula hospedeira, respectivamente. Esse estudo
esclareceu parte das discrepâncias em relação ao papel do remodelamento da
actina na invasão pelo T. cruzi, ao mostrar que a dependência ou não da actina
pode ser determinada pela molécula de superfície que é utilizada pelo parasita
na invasão. Após a interação da molécula com seu receptor na célula-alvo,
diferentes vias de sinalização levando à reorganização da actina são ativadas.
Em estudos anteriores, Schenkman e Mortara (1992) observaram que a
entrada das formas metacíclicas não é afetada pelo tratamento das lulas
com citocalasina D, enquanto Osuna et al. (1993) observaram que o processo
de invasão dessas formas é significativamente inibido por latrunculina B, droga
que desorganiza o citoesqueleto de actina. Resultados contrastantes também
foram encontrados em experimentos com formas TCT. Alguns estudos
mostraram que a internalização dessas formas é independente do
citoesqueleto de actina da lula hospedeira, pois o tratamento das células
com citocalasina D, droga que despolimeriza a F-actina, não inibe ou aumenta
a invasão dos parasitas (Schenkman et al., 1991b, Tardieux et al., 1994).
Outros estudos mostraram que a citocalasina D tem efeito inibitório na invasão
de diversos tipos celulares, incluindo fibroblastos, mioblastos, cardiomiócitos e
macrófagos (Barbosa & Meirelles, 1995, Rosestolato et al., 2002).
23
Objetivos do Estudo
Capítulo 1 Bases moleculares da avirulência do clone CL-14 do
Trypanosoma cruzi
Investigar as causas da aviruncia das formas metacíclicas do clone
CL-14, em comparação com a cepa CL, com foco no papel da gp82 de
superfície.
Capítulo 2 Expressão e localização celular de moléculas da família gp82
em formas metacíclicas do Trypanosoma cruzi
Caracterizar o repertório das moléculas gp82 expressas nas formas
metacíclicas através da análise da expressão e localização celular nas duas
cepas filogeneticamente distintas, G e CL.
Capítulo 3 Indução de apoptose em lulas de melanoma pela gp82 em
sua forma recombinante
Investigar o efeito p-apoptótico da proteína gp82 do T. cruzi, em sua
forma recombinante (J18), sobre as células do melanoma Tm5 e sobre a
linhagem de melanócitos melan-a.
24
Capítulo 1
Bases moleculares da avirulência do clone
CL-14 do Trypanosoma cruzi
Artigo:
“Molecular basis of non-virulence of Trypanosoma cruzi clone CL-14”
Atayde, V., Neira, I., Cortez, M., Ferreira, D., Freymüller, E., Yoshida, N.
International J. Parasitology, 2004, 34: 851-860
25
As formas metacíclicas da cepa CL (Brener & Chiari, 1963), que utilizam
a gp82 de superfície na invasão celular, são altamente infectivas in vitro e in
vivo (Ruiz et al., 1998, Neira et al., 2003). Em contraste, as formas metacíclicas
do clone CL-14, derivado da cepa CL (Chiari, 1981), não produzem infecção
patente mesmo quando inoculadas em camundongos recémnascidos (Lima et
al., 1990). No entanto, elicitam mecanismos imunológicos protetores contra
desafios letais com formas sangneas de cepas virulentas (Lima et al., 1990,
Paiva et al., 1999). É possível que a manutenção do clone CL-14 em culturas
axênicas por longos períodos tenha contribdo para a sua atenuação (Lima et
al., 1990).
Nosso principal objetivo foi determinar os fatores associados com a
baixa infectividade do clone CL-14. A invasão celular in vitro com formas
metacíclicas desse isolado é possível e foi utilizada em nosso estudo, sempre
em comparação com a cepa parental, CL (Fig.15). Os resultados mostraram
que a reduzida capacidade de invasão do clone CL-14 esassociada com a
baixa expressão da gp82 em sua superfície e que, aparentemente, utiliza a
gp35/50 para a sua internalização. Esses dados reforçam a importância da
gp82 da superfície das formas metacíclicas no processo de invasão celular.
CL-14
CL
Figura 15. Invasão celular in vitro por formas metacíclicas da cepa CL e do clone
CL-14. Notar a reduzida quantidade de parasitas do clone CL-14 (Giemsa, 1000x).
26
Experimentos complementares:
Imagens de lulas HeLa contendo amastigotas da cepa CL e do clone
CL-14 do T. cruzi
A figura 1C do artigo mostra que o clone CL-14 não é capaz de se
replicar intracelularmente. Em ensaios mantidos por 72 horas após a invasão
com as formas metacíclicas, quase não foram encontrados amastigotas
intracelulares do clone CL-14. No caso da cepa CL, muitas células foram
encontradas contendo um grande número de amastigotas. Na figura 16,
apresentamos imagens que ilustram essa observação.
Figura 16. Replicão intracelular da cepa CL e do clone CL-14 (Giemsa, 1000x).
Níveis steady-state” dos transcritos gp82 da cepa CL e do clone CL-14
do T. cruzi
Os níveis “steady-statedos transcritos gp82 de formas metacíclicas da
cepa CL e do clone CL-14 foram analisados por “northern blot. O RNA total
extraído dos parasitas foi hibridizado com a sonda J18, produzida a partir de
um fragmento de cDNA que contém a fase de leitura aberta do gene gp82,
mesma sonda utilizada para as análises de “southern blot” do artigo. Em
ambos, foi revelado um fragmento de aproximadamente 2.4 Kb (Fig.17). Esse
27
resultado corrobora os dados do artigo que mostram que o clone CL-14, apesar
de não expressar a gp82 na superfície, contém a molécula intracelular.
Figura 17. veis steady-state dos transcritos gp82 da cepa CL e do clone CL-
14 do T. cruzi. (A) Formas metacíclicas (1 x 10
8
) foram ressuspendidas em 1 ml de
Trizol (Invitrogen). As a homogeneização, foram adicionados 200 l de clorormio e
a amostra foi centrifugada a 10.800 rpm, 15 min. À fase aquosa, foram adicionados
500 l de isopropanol. Em seguida, a amostra foi centrifugada, lavada com etanol
70% e depois de seca, ressuspendida em água com dietilpirocarbonato a 0.05%.
Quantidades de RNA (~8 g) extraído foram desnaturadas a 70
o
C por 15 min em
tampão de amostra (50 l de MOPS 10X (MOPS 41.6g/l, acetato de sódio 6.76g/l,
EDTA 2.72g/l, pH 7.0), 90 l de formaldeído 37%, 250 l de formamida e 2 l de
solução de azul de bromofenol 0.01% em MOPS 1X). Após a desnaturação, as
amostras foram mantidas em gelo e momentos antes da aplicação no gel (1%
agarose, 10% formaldeído em MOPS 1X) foram adicionados 1.5 l de brometo de
edio. (B) Para alise por “northern blot”, após a eletroforese, o RNA do gel foi
transferido à membrana e em seguida ligado covalentemente ao nylon através da
exposição à luz UV (150 mJ), no sistema GS Gene Linker UV Chamber (Bio-Rad). Em
seguida, a membrana foi hibridizada com a sonda J18. Sonda α-tubulina foi utilizada
como controle positivo do experimento. As a hibridização, a membrana foi exposta
em filme de raio-X por 72 horas a -70
o
C. Os meros à esquerda indicam tamanhos
moleculares.
28
Capítulo 2
Expreso e localizão celular de moléculas
da família gp82 em formas metacíclicas
do Trypanosoma cruzi
Artigo:
“Expression and cellular localization of molecules of the gp82 family
in Trypanosoma cruzi metacyclic trypomastigotes”
Atayde, V., Cortez, M., Souza, R., Franco da Silveira, J., Yoshida, N.
Infection and Immunity, 2007, 75: 3264-3270
29
Pelo menos 50% do genoma do T. cruzi é composto por seqüências
repetidas, que podem ser retrotransposons, repetições subteloméricas ou
seqüências pertencentes às famílias que codificam proteínas de superfície.
Essas famílias multigênicas (MASP, trans-sialidases, mucinas e proteases
gp63), representam aproximadamente 18% dos 22.570 genes codificadores de
proteínas do parasita (El-Sayed et al., 2005).
O T. cruzi apresenta em seu genoma aproximadamente 60 cópias de
seqüências homólogas ao gene gp82, formando a família gênica gp82 (Araya
et al., 1994). Essa família é parte da grande família gp85/trans-sialidase,
composta por 1430 genes, representando aproximadamente 6% dos genes
codificadores do parasita (El-Sayed et al., 2005). Através da análise da
estrutura e expressão de genes da família gp82, Songthamwat et al. (2007)
identificaram três subfamílias nicas gp82, com expressão de RNA
mensageiro (mRNA) estágio-específica nas formas metacíclicas.
A gp82 apresenta 53.3% de identidade de aminoácidos com a gp85,
molécula envolvida na invasão das formas TCT que conm sítios de ligação à
laminina e à citoqueratina 18 (Giordano et al., 1999, Magdesian et al., 2001).
As regiões de homologia incluem os motivos Asp (SXDXGXTW), descritos nas
neuraminidases bacterianas, e o motivo VTV (VTVXNVFLYNR), específico do
T. cruzi, ambos característicos da família gp85/trans-sialidase (Crennell et al.,
1993, Cross & Takle, 1993, Frasch, 2000) (Fig.18). Apenas 25% da família
conservam o motivo VTV, provavelmente devido à forte pressão seletiva sobre
seus membros, que são alvos do sistema imune tanto por respostas humorais
como celulares (El-Sayed et al., 2005, Tzelepis et al., 2008). No extremo final
das duas seqüências há uma região hidrofóbica que é sinal para adição de
âncora GPI (Araya et al., 1994, Giordano et al., 1999) (Fig.18).
Duas moléculas gp82 de superfície, reativas com o MAb 3F6, foram
caracterizadas a partir de clones de cDNA das cepas G (J18) e CL (R31). J18 e
R31 possuem identidade de 97,9% entre suas seqüências peptídicas (Araya et
al., 1994, Ruiz, 1998, Yoshida et al., 2006).
30
Figura 18. Alinhamento das seências de aminoácidos da gp85 (clone Tc85-11,
GeneBank AAD13347.1) e da gp82 (clone J18, GeneBank AAA21303). Preto:
representa 100% de identidade, amarelo: motivos Asp, rosa: motivo VTV, verde: sinal
de âncora GPI. No caso da gp82, o motivo VTV é subterminal e é detectado um
terceiro motivo Asp degenerado.
O clone J18 contém 2.139 pb e sua fase de leitura aberta (ORF) vai do
nucleotídeo 231 (ATG) ao 1779 (TGA), contendo 1548 nucleotídeos (Araya et
al., 1994). Codifica um polipeptídeo de 516 aminoácidos com massa molecular
de 55,6 KDa, menor do que a massa de 70 kDa do polipeptídeo sem os
oligossacarídeos N-ligados, precursor da gp82 (Ramirez et al., 1993). Para
tentar explicar essa discrepância, uma análise computacional apontou a
31
existência de diversos sítios de miristoilação, além dos sítios de N-glicosilação.
O tamanho dos mRNA transcritos a partir do gene gp82 foram compatíveis com
o da ORF clonada (Araya et al., 1994). Já o clone R31 contém 2.150 pb e sua
ORF vai do nucleotídeo 190 (ATG) ao 1786 (TGA), contendo 1596
nucleotídeos (Ruiz, 1998). Anterior à ORF há ainda 77 aminoácidos em fase de
leitura, sugerindo que a metionina encontrada pode não ser a primeira e que,
parte da região N-terminal da seência foi perdida. Sendo assim, R31 pode
ser considerado outro membro da família gp82 de maior tamanho em
comparação ao J18 (Ruiz, 1998).
Utilizando peptídeos sintéticos baseados na seqüência do clone J18,
Manque et al. (2000) mapearam o epítopo do MAb 3F6 (P3, 244-263) e o sítio
de adesão à célula hospedeira da gp82 de superfície, ambos presentes na
região do domínio central da molécula. O sítio de adesão celular é composto
pela justaposição de duas seqüências de aminoácidos separadas por outra
altamente hidrofóbica, e sobrepõe-se parcialmente ao epítopo do MAb 3F6 (P4,
254-273 e P8, 294-313) (Fig.19). Isso explica o fato desse anticorpo inibir a
invasão in vitro e in vivo de formas metacíclicas de cepas que utilizam
preferencialmente a gp82 (Ramirez et al., 1993). J18 e R31 possuem 100% de
identidade na seqüência de aminoácidos da região do sitio de adesão à lula
hospedeira (Yoshida, 2006).
Figura 19. Representação esquemática da gp82 de superfície de formas
metaclicas do T. cruzi (clone J18).
32
Baseando-se na existência da família gênica gp82, nosso objetivo foi
caracterizar o repertório das moléculas gp82 expressas nas formas
metacíclicas, através da análise da expressão e localização celular nas cepas
G e CL. Para isso, clonamos um novo membro da família que, em contraste
com J18, não possui o epítopo do MAb 3F6, que caracteriza as moléculas gp82
de superfície. Os dados desse estudo mostraram que moléculas pertencentes
à família gp82 não são expressas somente na superfície das formas
metacíclicas, sugerindo outras funções além da adesão celular. Além disso,
observamos que a cepa G, juntamente com outras cepas do grupo T. cruzi I,
possui um conjunto de moléculas gp82 intraflagelares.
33
Experimentos complementares:
Análise da expressão de moléculas da família gp82 em formas
metacíclicas das cepas G e CL
Utilizando os anticorpos policlonais contra as proteínas recombinantes
J18 e C03, contendo ou não o epítopo do MAb 3F6, respectivamente,
analisamos os extratos de formas metacíclicas das cepas G e CL por
immunoblot. Na cepa CL, uma única banda foi detectada pelo MAb 3F6,
enquanto os anticorpos policlonais anti-J18 e anti-C03 reconheceram bandas
duplas (Fig.20A). Em immunoblot bidimensional, as moléculas reativas com o
MAb 3F6 focalizaram em pH 4.6 a 5.1 e as moléculas reativas com os
anticorpos anti-J18 e anti-C03, em pH 4.6 a 5.4 (Fig.20B). Na cepa G, ao
menos duas bandas foram reconhecidas por cada anticorpo (Fig.21A). Em
immunoblot bidimensional, as proteínas da família gp82 reativas com o MAb
3F6 ou com o anticorpo anti-J18, focalizaram em pH 4.9 a 5.7, e aquelas
reativas com o anticorpo anti-C03 focalizaram em pH 4.9 a 5.5 (Fig.21B).
Localização celular de moléculas gp82 reconhecidas pelo anticorpo
policlonal anti-C03 em cepas do grupo T. cruzi I ou II
Um resultado inesperado de nossos estudos foi a reação preferencial do
anticorpo anti-C03 com componentes intraflagelares da cepa G, o que não foi
observado na cepa CL (Fig. 4 e 5 do artigo). Como as cepas G e CL pertencem
a grupos genéticos divergentes, procuramos determinar se a localização
flagelar de proteínas gp82 reativas com o anticorpo anti-C03 seria uma
característica de cepas do grupo T. cruzi I, examinando duas outras cepas
deste grupo. Formas metacíclicas das cepas 262 (José-IMT) e 1161 (cedidas
pela Dra. Marta Teixeira, USP), permeabilizadas e incubadas com o anticorpo
anti-C03, apresentaram o mesmo perfil da cepa G, com marcação
predominantemente flagelar (Fig.22). Examinamos tamm formas
metacíclicas de outras duas cepas do grupo T. cruzi II, SC (Covarrubias et al.,
34
2007) e Y (Silva e Nussenzweig, 1953), quanto à reação com o anticorpo
policlonal anti-C03. Observamos padrão similar ao da cepa CL, com marcação
heterogênea no corpo do parasita (Fig.22). Conclmos que a localização
preferencial das proteínas gp82 no flagelo pode ser grupo-dependente.
Figura 20. Expressão de moléculas da família gp82 em formas metacíclicas da
cepa CL do T. cruzi. (A) Extratos dos parasitas foram analisados por “immunoblot”
com os anticorpos indicados. (B) “Immunoblot” bidimensionais de extratos de formas
metaclicas mostrando proteínas separadas por focalização isoelétrica, seguida de
separação por SDS-PAGE e reação com os anticorpos indicados. Os números à
esquerda indicam as massas moleculares (kDa).
35
Figura 21. Expressão de moléculas da família gp82 em formas metacíclicas da
cepa G do T. cruzi. (A) Extratos dos parasitas foram analisados por “immunoblot” com
os anticorpos indicados. (B) Immunoblot bidimensionais de extratos de formas
metaclicas mostrando proteínas separadas por focalização isoelétrica, seguida de
separação por SDS-PAGE e reação com os anticorpos indicados. Os meros à
esquerda indicam as massas moleculares (kDa).
36
Figura 22. Localização celular de moléculas da família gp82 reconhecidas pelo
anticorpo policlonal anti-C03 em cepas do grupo T. cruzi I (262 e 1161) ou II (SC e
Y). Tripomastigotas metacíclicos vivos ou permeabilizados foram incubados com o
anticorpo anti-C03 e processados para visualização em microscópio de fluorescência.
Notar a localização predominantemente flagelar nas cepas do grupo T. cruzi I
permeabilizadas.
37
Expressão de moléculas da família gp82 em outras formas do T. cruzi
Os genes codificadores de proteínas dos tripanossomatídeos estão
organizados em unidades transcricionais policistrônicas, como ocorre nos
procariontes (Ullu et al., 1996). Sendo assim, mRNA de proteínas estágio-
específicas podem ser transcritos nas outras formas do parasita. Em análises
de transcrição do gene gp82, foram detectadas pequenas quantidades de
mRNA em epimastigotas, amastigotas e tripomastigotas (Carmo et al., 1999,
Manque et al., 2003, Songthamwat et al., 2007).
Restava saber se os transcritos gp82, presentes nas outras formas do
parasita além das formas metacíclicas, são traduzidos em proteínas. Para
responder a essa pergunta, analisamos a expressão de moléculas da família
gp82 nas formas epimastigotas, amastigotas e TCT, das cepas G e CL, por
immunoblot revelados com os anticorpos indicados (Fig.23).
Na cepa CL, o MAb 3F6 reagiu com moléculas gp82 somente das
formas metacíclicas, corroborando dados anteriores. O anticorpo policlonal
anti-J18, além das formas metacíclicas, reconheceu bandas de 60-100 kDa das
formas TCT, uma banda de ~110 kDa de amastigotas intracelulares e bandas
de ~70 e 82 kDa de amastigotas extracelulares. O anticorpo policlonal anti-
C03, além das formas metacíclicas, reconheceu bandas das formas TCT de
~82 e 95 kDa e não reconheceu nenhuma molécula de amastigotas (Fig.23A).
Na cepa G, o MAb 3F6 também reagiu com moléculas gp82 somente
das formas metacíclicas. O anticorpo policlonal anti-J18, além das formas
metacíclicas, reconheceu bandas de 67-100 kDa das formas TCT e bandas de
~70 e 82 kDa de amastigotas extracelulares. O anticorpo policlonal anti-C03,
além das formas metacíclicas, reconheceu uma banda de ~95 kDa das formas
TCT e uma banda de ~67 kDa de formas amastigotas extracelulares (Fig.23B).
Os dados mostram que moléculas da família gp82 sem o epítopo do
MAb 3F6 o expressas em todas as formas do T. cruzi, exceto em
epimastigotas. Porém, a expressão é diferencial entre as cepas G e CL. Além
disso, os anticorpos policlonais anti-J18 e anti-C03, apesar de apresentarem
38
reatividade cruzada (Fig.3 do artigo), reagem com diferentes membros da
família gp82 nas formas do parasita analisadas.
A Cepa CL
B Cepa G
Figura 23. Expressão de moléculas da família gp82 em diferentes formas do T.
cruzi. Extratos dos parasitas foram analisados por immunoblot” com os anticorpos
indicados. (A) Cepa CL. (B) Cepa G. E, epimastigotas, M, formas metacíclicas, T,
tripomastigotas de cultura de tecido, AI, amastigotas intracelulares, AE, amastigotas
extracelulares. Osmeros à esquerda indicam as massas moleculares (kDa).
39
Capítulo 3
Indução de apoptose emlulas de melanoma pela
gp82 em sua forma recombinante
Artigo:
“A recombinant protein based on Trypanosoma cruzi surface molecule
gp82 induces apoptotic cell death in melanoma cells”
Atayde, V., Jasiulionis, M.G., Cortez, M., Yoshida, N.
Melanoma Research, 2008, 18: 172-183
40
O melanoma inicia-se a partir da transformação maligna do melanócito,
célula produtora de pigmento encontrada na lâmina basal da epiderme. Essa
transformação ocorre devido a diversos fatores genéticos, epigenéticos e
ambientais (Satyamoorthy & Herlyn, 2002, Correa et al., 2005). Normalmente,
os melanócitos sintetizam melanina e a transferem em grânulos aos
queratinócitos subjacentes, que a armazenam. Essa pigmentação protege
contra os danos causados pela radiação ultravioleta solar (Ueda & Richmond,
2006) (Fig.24).
Figura 24. Melanócitos na lâmina basal da epiderme. Em destaque es a
transferência de gnulos de melanina aos queratinócitos (training.seer.cancer.gov).
A linhagem de melanócitos melan-a, derivada de camundongo C57BL,
possui características de melanócitos normais, exceto senescência e resposta
proliferativa ao PMA (phorbol-12-myristate-13-acetate), ativador da PKC
(Bennett et al., 1987, Furstenberger et al., 1981). A partir de melan-a, foi
estabelecida a linhagem de melanoma Tm5, que apresenta crescimento
independente de PMA, menores tempos de duplicação e diferente adesividade
41
a componentes da matriz extracelular. Além disso, Tm5 é altamente
tumorigênica in vivo (Oba-Shinjo et al., 2006).
A incapacidade das células em entrar em apoptose (Fig.25) contribui
para a patogênese de diversos nceres, incluindo o melanoma (Hanahan &
Weinberg, 2000, Herlyn et al., 2002).
Figura 25. Vias clássicas de indão da apoptose. Na via extrínseca, é ativada a
caspase-8. Na via intrínseca, é ativada a caspase-9. IAP, c-FLIP, Bcl-2, HSP70 e
HSP90 são inibidores endógenos da apoptose. Bax e p53 são indutores (adaptado de
Heussler et al., 2001).
42
A morte celular por apoptose é induzida por duas vias principais: uma
extrínseca, que envolve os membros da família de receptores do TNF (fator de
necrose tumoral) com recrutamento e ativação da caspase-8, e outra
intrínseca, através da liberação do citocromo C mitocondrial e sua interação
com Apaf-1 no citoplasma, com ativação da caspase-9 (Fig.25). As duas vias,
que possuem conexões entre si (Taylor et al., 2008), convergem para ativar as
caspases efetoras 3, 6 e 7, responveis pelas alterações específicas da
apoptose: condensação e clivagem do DNA, exposição da fosfatidilserina na
porção externa da membrana plasmática e fragmentação celular em corpos
apoptóticos (Fig.26), dentre outras (Heussler et al., 2001, Fulda & Debatin,
2006). A liberação do citocromo C é controlada pela família bcl-2 de proteínas,
que tem sua transcrição regulada pelo fator NF-κB, constitutivamente ativo em
células de melanoma (Richmond, 2002, Fecker et al., 2005, Amiri and
Richmond, 2005, Ueda and Richmond, 2006).
Figura 26. Corpos apoptóticos observados por microscopia eletrônica de
varredura. Os corpos apopticos o permitem o extravasamento do conteúdo
intracelular no tecido, protegendo-o de inflamação (mun.ca).
Durante a apoptose, o citoesqueleto de actina, responvel pelo formato,
motilidade celular e transdução de sinais exteriores (Ho et al., 2008), é alvo das
caspases efetoras. Mudanças na dinâmica da actina filamentosa (F-actina), por
sua vez, induzem a ativação das caspases (Leadsham & Gourlay, 2008).
43
O aumento do turnover” da F-actina promove a longevidade celular,
enquanto que a diminuição pode induzir a apoptose (Gourlay & Ayscough,
2005, Thomas et al., 2006). Drogas que alteram a dinâmica da F-actina, como
a citocalasina D, ativam a apoptose em lulas humanas e murinas através de
uma via caspase-3-dependente (Posey & Bierer, 1999, Yamazaki et al., 2000,
Odaka et al., 2000, Boldogh & Pon, 2006). Ainda não se sabe ao certo por qual
mecanismo a caspase-3 é ativada nesses casos. A proteína recombinante J18,
baseada na molécula de superfície gp82 de formas metacíclicas do T. cruzi
(Araya et al., 1994), induz despolimerização do citoesqueleto de actina, efeito
análogo ao da citocalasina D (Cortez et al., 2006b) (Fig.27). Esse efeito é
dependente da liberação de Ca
+2
induzida por J18 na lula hospedeira (Ruiz
et al., 1998, Cortez et al., 2006b).
Figura 27. Diminuição das fibras de actina em células HeLa tratadas com J18 ou
citocalasina D por 30 minutos. Verde: actina, azul: núcleos (Cortez et al., 2006b).
As interações entre a actina e as mitocôndrias, que agem como
sensores da diminuição do turnover” da actina, são centrais na regulação da
apoptose pelo citoesqueleto. (Gourlay and Ayscough, 2005, Leadsham &
Gourlay, 2008). Quando a actina é estabilizada, um canal iônico voltagem-
depentente (VDAC) da mitocôndria é aberto, permitindo a liberação de
moléculas apoptogênicas como citocromo C, AIF e smac (Gourlay & Ayscough,
44
2005). Além disso, alguns membros da família bcl-2 podem interagir
diretamente com VDAC, determinando a permeabilidade da membrana
mitocondrial (Tsujimoto, 2002). Transições na permeabilidade alteram o
potencial transmembrânico, um dos primeiros sinais apoptóticos de células de
mamíferos (Simeonova et al., 2004) (Fig.28).
Figura 28. Regulação de VDAC pela actina. A redução na dinâmica da actina pode
levar à redução no potencial mitocondrial e indução da apoptose após a liberação de
fatores apoptogênicos (adaptado de Gourlay & Ayscough, 2005).
Muitas evincias experimentais mostram que infecções com diversos
patógenos, como Salmonella typhimurium (Pawelek et al., 1997), Toxoplasma
gondii (Hunter et al., 2001) ou Trypanosoma cruzi (Oliveira et al., 2001), podem
retardar ou inibir o crescimento de tumores malignos, inclusive dos melanomas,
em animais e humanos. Para explicar o efeito antitumoral das infecções pelo T.
cruzi, alguns pesquisadores propuseram a existência de uma toxina secretada
pelos parasitas, outros a ocorrência de resposta imune contra antígenos das
células tumorais (Cabral, 2000).
45
Pouco se sabe sobre a produção de fatores solúveis ou secretados por
patógenos que podem agir nas lulas tumorais, causando sua morte. A
proteína azurina, de Pseudomonas aeruginosa, entra em lulas de melanoma
humano e induz apoptose devido ao aumento dos níveis intracelulares de p53
(Yamada et al., 2002). Uma fração rica em lipídeos, isolada de promastigotas
de Leishmania donovani, induz apoptose em melanomas de camundongo e
humano (Ratha et al., 2006).
Evidências sobre as propriedades antitumorais do T. cruzi mostram o
efeito citotóxico direto de lisados dos parasitas sobre lulas tumorais in vitro
(Kallinikova et al., 2001, Sheklakova et al., 2003), efeito que deve existir devido
ao potencial pró-apoptótico que é visto em algumas moléculas do T. cruzi
contra células do sistema imune. Uma dessas moléculas é a transialidase
(Freire-de-Lima et al., 1998, Borges et al., 2005, Mucci et al., 2006).
Nosso principal objetivo foi investigar o efeito pró-apoptótico da proteína
gp82 recombinante (J18) sobre células do melanoma Tm5, comparado ao
efeito sobre a linhagem de melanócitos melan-a. Os resultados mostraram que
o tratamento com J18 induz a despolimerização da F-actina somente nas
células do melanoma Tm5, levando à indução da apoptose através da ativação
da caspase-3. Sinais apoptóticos específicos como encolhimento nuclear e
citoplasmático, exposição da fosfatidilserina e fragmentação do DNA nuclear,
foram observados. Também notamos a perda do potencial transmembrânico e
a localização citoplasmática de NF-kB decorrentes do efeito de J18 no
citoesqueleto de actina.
46
Experimentos complementares:
Análise da proteína recombinante J18:
Uma questão em relação à proteína recombinante J18 que não havia sido
ainda determinada é como ela se apresenta na forma purificada, se como
monômero, dímero ou aglomerado protéico. Para responder a essa pergunta,
purificamos uma preparação de J18 por gel filtração em sistema FPLC e
obtivemos três principais picos correspondentes a: proteínas maiores de 160
KDa representando aglomerados de J18, proteínas de aproximadamente 160
kDa representando dímeros de J18, proteínas de aproximadamente 80 kDa
representando monômeros de J18 (Fig.29A). Submetemos as frações 11 a 43 à
SDS-PAGE em condições desnaturantes, seguido de “immunoblot” revelado
com o MAb 3F6. Como mostrado na Figura 29B, bandas de aproximadamente
80 kDa foram detectadas nas frações 14 a 30 e uma maior quantidade de
proteína foi detectada nas frações 20 a 24, representando aglomerados e
meros de J18 (>160 kDa e ~160 kDa). Selecionamos algumas frações
representantes dos diferentes picos da gel filtração e submetemos à eletroforese
não-desnaturante seguida de immunoblot e revelação com o MAb 3F6. Foram
detectadas bandas maiores de 80 kDa, atingindo até ~230 kDa, principalmente
nas frações 21 e 24, representantes dos picos de proteínas maiores de 160 kDa
e de ~160 kDa (Fig.30).
O efeito inibitório de J18 sobre a invasão celular de formas metacíclicas é
conhecido. Para comparar com o efeito de J18 na sua forma monomérica,
dimérica ou oligomérica, realizamos experimentos com as seguintes
preparações: frações 11, 21, 24 e 27, a preparação completa de J18, o
sobrenadante e o “pellet” de uma preparação de J18 centrifugada a 13.000 rpm
por 15 minutos. A inibição da invasão foi semelhante (~40%) para todas as
amostras, exceto para a fração 11, que foi utilizada como controle (Fig.31).
47
Figura 29. Purificão da proteína recombinante J18 por gel filtração em sistema
FPLC. (A) Uma preparação de J18 foi purificada em coluna cromatográfica de gel
filtração Superdex HR200 em sistema FPLC, como descrito (Amino et al., 2002). (B)
As frações de 0.5 ml coletadas foram submetidas à SDS-PAGE e “immunoblot
revelado com o MAb 3F6. Notar o reconhecimento das frações 14 a 30 pelo MAb 3F6,
de acordo com os picos do gráfico.
48
Figura 30. “Immunoblot” de preparações da proteína recombinante J18
submetidas à SDS-PAGE não-desnaturante. As frações coletadas da gel filtração
(11, 21, 24 e 27), a proteína total (J18), a proteína centrifugada (C), ou o “pellet” (P)
foram submetidos à SDS-PAGE o-desnaturante, transferidos à membrana de
nitrocelulose e revelados com o MAb 3F6. Notar o reconhecimento de bandas maiores
de 80 kDa principalmente nas frações 21 (> 160 kDa) e 24 (~160 kDa). Os números à
esquerda representam as massas moleculares (kDa).
Figura 31. Invasão celular de formas metacíclicas da cepa CL na presea de
preparações da proteína recombinante J18. Formas metaclicas da cepa CL foram
incubadas por 1 hora com células HeLa (10: 1) na presença das frações coletadas da
gel filtração (11, 21, 24 e 27), da proteína total (J18), da proteína centrifugada (C), ou
do “pellet da centrifugação (P), na concentração final de 40 µg/ml. O total de parasitas
em 100 células foi estimado contando-se lamínulas coradas com Giemsa em
duplicatas (500 células por lamínula). Experimento representativo de três
experimentos independentes.
49
Y*
Com esses resultados, mostramos que uma preparação de J18 contém
proteínas com diferentes níveis de agregação e que essas diferentes formas têm
a mesma capacidade de interagir com a célula hospedeira, competindo com os
parasitas no momento da invasão. Dessa maneira, obtivemos mais informações
em relação ao preparado (J18) utilizado no tratamento das lulas melan-a e
Tm5.
Imagens de tumores em camundongos infectados com T. cruzi
Camundongos C57BL/6 foram infectados com formas metacíclicas de
duas cepas de T. cruzi expressando ou não a gp82 na superfície, cepas G e Y*,
respectivamente. Três semanas após a infecção, os camundongos receberam
células do melanoma Tm5 (2 x 10
5
) por via subcutânea na região dorsal e o
desenvolvimento dos tumores foi acompanhado. Os volumes tumorais foram
significativamente menores nos animais infectados com a cepa G, comparados
aos volumes tumorais dos camundongos infectados com a cepa Y* ou dos
camundongos não-infectados, como mostrado na figura 9 do artigo (Fig.32).
Figura 32. Desenvolvimento tumoral (20 dias pós-inoculação de células Tm5) em
camundongos infectados com T. cruzi.
50
Discussão
Nosso principal objetivo foi caracterizar a família gp82 de proteínas do T.
cruzi. Até o início desse projeto, somente a gp82 de superfície, identificada pelo
MAb 3F6 e envolvida na invasão celular das formas metacíclicas, havia sido
descrita. Mostramos que membros da família gp82 são expressos
intracelularmente e que alguns têm localização diferenciada em parasitas dos
grupos T. cruzi I e II. Ao contrário das glicoproteínas gp82 reativas com o MAb
3F6, específicas das formas metacíclicas (Teixeira & Yoshida, 1986), outras
moléculas da família foram detectadas nas formas TCT e amastigotas.
Descrevemos ainda o efeito antitumoral da gp82 de superfície em sua forma
recombinante, propriedade nunca antes descrita para outro componente do T.
cruzi molecularmente definido.
Na primeira parte do estudo, investigamos as bases da avirulência das
formas metacíclicas do clone CL-14, derivado da cepa CL. Descobrimos que a
baixa infectividade desses parasitas está associada com a reduzida expressão
da gp82 na superfície, reforçando o papel central da molécula na infecção pelo
T. cruzi. Além de invadir cerca de quatro vezes menos do que a cepa CL in
vitro, as formas amastigotas do clone CL-14 não foram capazes de se replicar
em células HeLa mantidas a 37
o
C. Esse resultado está de acordo com achados
anteriores onde foi mostrado que a temperatura permissiva para a
diferenciação do clone CL-14 em células não-fagocíticas é de 33
o
C (Almeida-
de-Faria et al., 1999). A temperatura como um fator do hospedeiro na inibição
do desenvolvimento intracelular é compatível com nossa observação de que as
formas metacíclicas do clone CL-14 não produzem parasitemia patente em
camundongos infectados por via oral.
Uma vez que a infecção oral por formas metacíclicas da cepa CL é
extremamente eficiente e é dependente da gp82 de superfície (Neira et al.,
2003), a deficiência na expressão dessa molécula no clone CL-14 seria o único
fator responvel pela baixa infectividade, não fosse seu desenvolvimento
limitado pela temperatura. Na infecção oral, além de promover a invasão das
51
células da mucosa gástrica, foi descrito que a interão entre a gp82 e a
mucina é fundamental (Neira et al., 2003). No estudo de formas metacíclicas de
cepas do T. cruzi gp82-deficientes, observou-se que essas são pouco
infectivas em camundongos por via oral, embora a invasão celular in vitro seja
eficiente devido à expressão da glicoproteína gp30, que tem características
similares à gp82 (Cortez et al., 2003). Gp30 liga-se à célula hospedeira de
maneira receptor-dependente e induz sinal de Ca
+2
, propriedade associada à
invasão celular. Entretanto, ao contrário da gp82, a capacidade de ligação da
gp30 à mucina gástrica é reduzida (Cortez et al., 2003). No caso do clone CL-
14, as formas metacíclicas não expressam na superfície nem a gp82 nem a
gp30, e são pouco invasivas tanto in vivo como in vitro.
As formas metacíclicas do clone CL-14 e da cepa G assemelham-se
quanto à baixa infectividade e aparentemente prescindem da gp82 para invadir
a célula hospedeira. É possível que o clone CL-14, como a cepa G, faça uso da
gp35/50 para interagir com a lula hospedeira. Entretanto, a gp35/50 expressa
no clone CL-14 é uma variante da gp35/50 da cepa G, que reage com o MAb
2B10 mas não é reconhecida pelo MAb 10D8 (Mortara et al., 1992). Embora as
formas metacíclicas da cepa G expressem gp82 na superfície em níveis
comparáveis aos da cepa CL, a interação da gp82 com a célula hospedeira é
prejudicada pela gp90, molécula que modula negativamente a penetração do
parasita (Málaga e Yoshida, 2001). O clone CL-14 expressa baixos níveis de
uma molécula variante da gp90 da cepa G, reconhecida pelo MAb 5E7, mas
não pelo MAb 1G7 (Yoshida, 2006). Nesse caso, a gp90 provavelmente não
atua como modulador negativo da invasão.
Na segunda parte do estudo, analisamos a expressão e a localização de
moléculas da família gp82 em formas metacíclicas das cepas G e CL. Para
isso, isolamos um novo membro, que não contém o epítopo do MAb 3F6.
Quando alinhamos as seqüências de aminoácidos dos clones C03 (gp82) e Tc-
85 (gp85 das formas TCT), observamos 57,2% de identidade, indicando que o
novo clone é parte da família gp85/trans-sialidase. As duas seqüências
possuem semelhanças na região N-terminal, que é inexistente no clone J18
(gp82). J18 e Tc-85 apresentam 53,3% de identidade (Fig.33).
52
Figura 33. Alinhamento das seências de aminoácidos dos clones J18, C03 e
Tc85. J18 (AAA21303), C03 (EF445668) Tc85-11 (AAD13347.1). Preto: representa
100% de identidade, amarelo: motivos Asp, rosa: motivo VTV, verde: sinal de âncora
GPI, vermelho: regiões do epítopo do MAb 3F6 e do sítio de adesão celular do clone
J18, azul: sítio de ligação à laminina do clone Tc-85.
Um achado interessante desse trabalho foi a marcação
predominantemente flagelar do anticorpo policlonal anti-C03 em formas
metacíclicas de cepas do grupo T. cruzi I. O flagelo dos tripanossomatídeos é
único e multifuncional, atuando na motilidade, quimiotaxia, sinalização e
invasão celular. Muitas moléculas do T. cruzi associadas ao flagelo têm sido
descritas. Uma delas, cuja função ainda não se conhece, tem 160 kDa (FL-160)
53
e é codificada por uma família gênica que é parte da grande família gp85/trans-
sialidase. Essa protna é expressa em formas TCT, na membrana do bolso
flagelar, região de onde emerge o flagelo (Van Voorhis et al., 1993). Um
membro dessa família gênica, FL-160-CRP, é expresso na superfície dos
parasitas, mas não é expresso no flagelo. Na análise dos transcritos da FL-
160-CRP, variações nas seqüências foram associadas com a divisão das
cepas em grupos T. cruzi I e II (Mathieu-Dauet al., 2008). Se o anticorpo
anti-C03 pode ser usado como um marcador de cepas do grupo T. cruzi I,
ainda precisa ser estabelecido. A função dos membros da família gp82
expressos no flagelo é desconhecida.
Em comparação com o clone J18, os clones C03 e Tc-85 apresentam
diversas substituições de aminoácidos nas regiões do epítopo do MAb 3F6 e
do tio de adesão celular da gp82 (Fig.33). Talvez por isso a gp85 não tenha
sido reconhecida pelo anticorpo monoclonal em extratos de formas TCT das
cepas G ou CL, nos ensaios de “immunoblot”. Porém, como possuem diversos
epítopos comuns, os anticorpos policlonais anti-J18 e anti-C03 reagiram com
um conjunto de moléculas dessas formas, que possivelmente inclui a gp85
(Fig.23).
Em relação à adesão celular, foi descrito que a molécula de superfície
gp85 contém sítios de ligação à laminina e à citoqueratina 18 (CK18) (Giordano
et al., 1999, Magdesian et al., 2001), interações que podem promover ou
facilitar a invasão das formas TCT. Utilizando peptídeos sintéticos, Giordano et
al. (1999) mapearam o sítio de adesão do clone Tc-85 à laminina (peptídeo
TGETPLEPFGFCFGA). Dos 15 aminoácidos que compõem a região referente
ao peptídeo de ligação à laminina, o clone J18 possui apenas 5 em comum,
sugerindo que se essa molécula também interage com a laminina é através de
outro sítio (Fig.33). A Tc-85 interage com a CK18 através do motivo VTV, que
aumenta a invasão das formas TCT (Magdesian et al., 2007). O clone J18
possui esse motivo na região subterminal (Fig.33), e a sua relevância na
adesão do parasita à célula é discutível.
54
Outras moléculas das formas TCT, além da gp85, interagem direta ou
indiretamente com moléculas da matriz extracelular (ECM). A gp83, molécula
de superfície de 83 kDa, aumenta a expressão da laminina-γ1 em células
musculares humanas após a exposição à sua forma recombinante. A
superexpressão da laminina torna as células mais susceptíveis à infecção (Nde
et al., 2006). A gp83 tem 32,8% de identidade na seqüencia de aminoácidos
com o clone J18 e 32,3% com o clone C03. É possível que os anticorpos
policlonais anti-C03 e anti-J8 reconheçam epítopos da gp83 das formas TCT
das cepas G ou CL, como visto na figura 23. Foram tamm detectadas, nas
formas TCT, proteínas ligantes de heparina e heparan sulfato de
aproximadamente 59 e 65,8 kDa, e essa ligação está envolvida na invasão de
cardiomcitos. Uma delas pode ser a penetrina, previamente descrita, que
possui as mesmas características (Ortega-Barria & Pereira, 1991, Oliveira Jr. et
al., 2008). A serino-protease POP Tc80, hidrolisa colágeno dos tipos I e IV e
fibronectina, provavelmente como modo de facilitar a migração através da ECM
(Grellier et al., 2001, Bastos et al., 2005). Também foi mostrado que formas
TCT possuem um receptor de superfície para a fibronectina (Ouaissi et al.,
1984).
A ligação de moléculas do T. cruzi à laminina, heparan sulfato,
fibronectina e outros componentes, pode ser essencial na migração dos
parasitas através da ECM em direção às lulas-alvo. É importante lembrar
que, de maneira similar, a gp82 interage com a mucina gástrica, facilitando o
acesso das formas metacíclicas às lulas do epitélio gástrico durante a
infecção oral (Neira et al., 2003). Até o momento, o receptor celular da gp82 de
superfície não foi identificado. Além disso, a qual componente da ECM se liga a
gp82 também não é conhecido. Foi mostrado que formas TCT entram em
células epiteliais preferencialmente pela porção baso-lateral (Schenkman et al.,
1991c). Se isso ocorrer também com as formas metacíclicas, uma hipótese
seria que a gp82 contém sítios de interação com componentes da lâmina basal,
já que essas formas, quando iniciam a infecção no hospedeiro vertebrado,
invadem células epiteliais. A gp82 parece não interagir com colágeno tipo IV,
55
presente na lâmina basal (Ramirez et al., 1993). Resta saber se interage com a
laminina.
Na terceira parte do estudo, ocorreu-nos a iia de examinar a ação de
J18 sobre as células de melanoma, em função das propriedades dessa
molécula reveladas em trabalhos anteriores. Cortez et al. (2006b) observaram
que J18 induz a despolimerização da F-actina de células HeLa. Levantamos a
possibilidade de que o tratamento das células com J18 por tempo prolongado
levaria, em última instância, à morte celular. Nos diversos experimentos, J18
mostrou efeito seletivo sobre as células do melanoma Tm5, mas não sobre as
células parentais não tumorinicas, melan-a. Esse efeito não se deve à falta
do receptor para essa molécula nas lulas melan-a, pois J18 liga-se a estas
células do mesmo modo que às células Tm5. Uma possibilidade é que, após a
ligação de J18 ao seu receptor, vias de sinalização distintas sejam induzidas
nas duas linhagens, levando à desestruturação dos filamentos de actina
somente nas células de melanoma (Fig.34). Em células HeLa, foi mostrado que
a despolimerização da F-actina é dependente de sinal de Ca
+2
, deflagrado
após a ativação de uma via de sinalização ainda desconhecida (Cortez et al.,
2006b).
Segundo dados da literatura, a citocalasina D induz morte celular por
apoptose em células transformadas através da ativação da caspase-3
(Yamazaki et al., 2000) sendo estas mais susceptíveis do que células normais
(Odaka et al., 2000). O fato das lulas tumorais conterem menor quantidade
de F-actina do que lulas e tecidos normais (Verderame et al., 1980, Varani et
al, 1983, Rao et al., 1990), poderia explicar a diferença entre as
susceptibilidades de melan-a e Tm5 à J18. Como J18 tem ação similar à da
citocalasina D (Cortez et al., 2006b), analisamos o tipo de morte induzida por
esta molécula nas lulas Tm5 e encontramos diversas evincias de
apoptose. O primeiro evento apoptótico detectado foi a ativação da caspase-3,
provavelmente responvel pela ocorncia dos eventos subseqüentes
(Fig.34). Não observamos ativação das caspases 8 e 9. Algumas moléculas do
T. cruzi apresentam efeito contrário ao de J18. A transialidase inibe a apoptose
em células de Schwan humanas, através de uma via ativada pela PI-3 quinase
56
(Chuenkova et al., 2001). A cruzipaína, cisteína protease envolvida na invasão,
crescimento e diferenciação do T. cruzi (Murta et al., 1990, Paiva et al., 1998,
Meirelles et al., 1992, Scharfstein et al., 2000), quando inativada, promove a
sobrevivência de cardiomiócitos de culturas primárias de camundongos através
da indução da expressão de Bcl-2, molécula anti-apoptótica (Aoki et al., 2006).
Figura 34. Representação esquemática da via apoptótica ativada em lulas do
melanoma Tm5 pelo tratamento com J18. A ligação de J18 com seu receptor (J18R)
na célula Tm5 leva à desorganização do citoesqueleto de actina com ativação da
caspase-3, responsável pela exposição da fosfatidilserina (PS), perda do potencial da
membrana mitocondrial m), inibição da translocação de NF-κB ao cleo e
fragmentação do DNA nuclear, fenômenos apopticos. Entre parênteses estão os
tempos de tratamento.
57
Uma possibilidade que não foi confirmada em nossos estudos é a
ativação da caspase-3 nas células Tm5, via ativação da caspase-12. A caspase-
12, que está envolvida na indução da apoptose após estresse de retículo
endoplasmático (Liu et al., 1998, Rao et al., 2002, Xie et al., 2002, Hitomi et al.,
2003) poderia, em seguida, ativar a caspase-3 nas lulas Tm5. Outra molécula
que poderia estar envolvida na ativação da apoptose em células Tm5 é a
calpaína, cisteína protease citossólica também intimamente relacionada com a
via de estresse de retículo endoplasmático. Essa enzima é ativada por Ca
+2
e
tem como alvo a caspase-12, ativando-a (Vandenabeele et al., 2005).
Investigamos ainda o efeito tumoricida de J18 in vivo. O
desenvolvimento do melanoma Tm5 foi significativamente mais lento em
camundongos tratados com J18 na região tumoral. Eventualmente, esses
resultados poderiam ser melhorados com outras dosagens de J18 e/ou tempo
de tratamento. O bloqueio da translocação de NF-kB aocleo in vitro pode ser
também a causa da inibição do crescimento de células Tm5 in vivo. Esse
resultado contrasta com resultados recentes com a transialidase em sua forma
recombinante que, ao invés de inibir, ativa a translocação de NF-kB ao núcleo
de células endoteliais humanas. Esse resultado é tamm observado quando
as células o tratadas com lipopolissacarídeo (LPS) bacteriano (Dias et al.,
2008). Contudo, não podemos descartar a hitese de que o J18 tenha outra
ação in vivo, diferente da observada in vitro. Talvez o crescimento dos tumores
seja reduzido pela indução de alterações importantes no microambiente
tumoral que, por sua vez, eliminariam as células tumorais dos tecidos.
Quanto à proteção ao desenvolvimento tumoral pela infecção com o T.
cruzi, descrita em diversos trabalhos, confirmamos esses achados infectando
camundongos com a cepa G, que é pouco infectiva, porém expressa altos
veis de gp82 na superfície, e com a cepa Y*, uma amostra da cepa Y, isolada
inicialmente por Silva & Nussenzweig (1953) de um paciente na fase aguda da
doença de Chagas. Formas metacíclicas de Y* expressam somente a gp30 na
superfície. Apesar das formas metacíclicas existirem por pouco tempo no
hospedeiro vertebrado, observamos significativa redução tumoral em
camundongos pré-infectados com a cepa G. Em contraste, camundongos pré-
58
infectados com a cepa Y*, que não expressa a gp82 reativa com o MAb 3F6,
mas expressa outros membros da família gp82 reconhecidos pelo anticorpo
policlonal anti-J18, foram incapazes de conter o crescimento tumoral.
A associação entre esses dados e os dados referentes à gp82
recombinante (J18) não é clara. O que pode ser hipotetizado é que moléculas
da família gp85/trans-sialidase, que têm seqüências em comum com a gp82, e
que são expressas em tripomastigotas sanguíneos, exerçam o papel de reduzir
o desenvolvimento dos tumores. Para citar um exemplo, a trans-sialidase
presente na corrente sangnea, secretada pelos parasitas, possui a
característica de induzir a apoptose de células do sistema imune do hospedeiro
(Mucci et al., 2006). Outra molécula que poderia ser responvel pela redução
tumoral é a gp85 expressa pelas formas sangneas da cepa G, uma vez que
vimos que as duas moléculas, gp82 e gp85 possuem epítopos comuns. Seria
interessante investigar se a gp85 de fato possui efeito antitumoral como
observamos com a gp82 recombinante. Outra explicação para a redução
tumoral em camundongos infectados com a cepa G seria a de que membros da
família gp82 expressos por formas sanguíneas ou amastigotas, quando
liberados no meio extracelular ou na corrente sanguínea, pudessem agir nas
células tumorais. Os membros da família gp82 expressos por formas
metacíclicas da cepa Y* não teriam função antitumoral.
Segundo Cabral (2000), a redução dos tumores de camundongos
infectados com T. cruzi ocorreria pelo desenvolvimento de uma resposta auto-
imune contra as lulas tumorais, devido à lise daquelas infectadas com o
parasita. Essa resposta ocorreria devido ao extravasamento e exposição do
conteúdo celular no meio extracelular. Alguns estudos mostram a auto-
imunidade à miosina cardíaca desenvolvida em camundongos ou humanos
infectados com T. cruzi. Porém, não se sabe ao certo se é resultado da lise e
destruição celular, ou da reatividade cruzada entre antígenos do parasita e do
hospedeiro (Engman & Leon, 2002). Dessa maneira, existe a possibilidade de
ocorrer reatividade cruzada entre moléculas dos parasitas da cepa G e das
células tumorais.
59
Nossos estudos abrem perspectivas para o estudo das proteínas da
família gp82, principalmente dos membros que possam ter funções diferentes
da função da gp82 de superfície, como por exemplo, os flagelares. Em
trabalhos subseqüentes, esses membros devem ser caracterizados
funcionalmente.
60
Referências Bibliográficas
Acosta-Serrano, A., Schenkman, S., Yoshida, N., Mehlert, A.,
Richardson, J.M. & Ferguson, M.A., 1995. The lipid structure of the
glycosylphosphatidylinositol-anchored mucin-like sialic acid acceptors of
Trypanosoma cruzi changes during parasite differentiation from epimastigotas
to infective metacyclic trypomastigote forms. J. Biol. Chem., 270: 27244-27253.
Acosta-Serrano, A., Almeida, I.C., Freitas-Junior, L.H., Yoshida, N. &
Schenkman, S., 2001. The mucin-like glycoprotein super-family of
Trypanosoma cruzi: structure and biological roles. Mol. Biochem. Parasitol.,
114: 143-150.
Almeida-de-Faria, M., Freymüller, E., Colli, W. & Alves, M.J., 1999.
Trypanosoma cruzi: characterization of na intracellular epimastigote-like form.
Exp. Parasitol., 92: 263-274.
Amiri, K.I. & Richmond, A., 2005. Role of nuclear factor-kappa B in
melanoma. Cancer Metastasis Rev., 24: 301-313.
Andrade, S.G., Rassi, A., Magalhães, J.B., Ferriolli Filho, F. &
Luquetti, A.O., 1992. Specific chemotherapy of Chagas disease: a comparison
between the response in patients and experimental animals inoculated with the
same strains. Trans. Royal Soc. Trop. Med. Hyg., 86: 624-626.
Aoki, M. del P., Cano, R.C., Pellegrini, A.V., Tanos, T., Guiñazú, N.L.,
Coso, O.A. & Gea, S., 2006. Different signaling pathways are involved in
cardiomyocyte survival induced by a Trypanosoma cruzi glycoprotein. Microbes
Infect., 8: 1723-1731.
Aragão, M.B., 1983. Domiciliação de triatomíneos ou pré-adaptação à
antropofilia e à ornitofilia? Rev. Saúde Públ., 17: 51-55.
Araya, J. E., Cano, M.I., Yoshida, N. & Franco da Silveira, J., 1994.
Cloning and characterization of a gene for stage specific 82-kDa surface
antigen of metacyclic trypomastigotes of Trypanosoma cruzi. Mol. Biochem.
Parasitol., 65: 161-169.
Aufderheide, A.C., Salo, W., Madden, M., Streitz, J., Bulkestra, J.,
Guhl, F., Arriaza, B., Renier, C., Wittmers Jr., L.E., Fornaciari, G. & Alisson,
A., 2004. A 9000-years record of Chagas’ disease. Proc. Nat. Acad. Sci. USA,
101: 2034-2039.
61
Bahia, M.T., Tafuri, W.L., Caliari, M.V., Veloso, V.M., Carneiro, C.M.,
Coelho, G.L. & Lana, M., 2002. Comparison of Trypanosoma cruzi infection in
dogs inoculated with blood or metacyclic trypomastigotes of Berenice-62 and
Berenice-78 strains via intraperitoneal and conjuctival routes. Rev. Soc. Bras.
Med. Trop., 35: 339-345.
Baptista, C.S., ncio, R.Z.N., Abdala, S., Carranza, J.C.,
Westenberger, S.J., Silva, M.N., Pereira, C.A.B., Galvão, L.M.C., Gontijo,
E.D., Chiari, E., Sturn, N. & Zingales, B., 2006. Differential transcription
profiles in Trypanosoma cruzi associated with clinical forms of Chagas’ disease:
Maxicircle NADH dehydrogenase subunit 7 gene truncation in asymptomatic
patient isolates. Mol. Biochem. Parasitol., 150: 236-248.
Barbosa, H.S. & Meirelles, M.N., 1995. Evidence of participation of
cytoskeleton of heart muscle cells during the invasion of Trypanosoma cruzi.
Cell. Struct. Funct., 20: 275-284.
Bastos, I.M., Grellier, P., Martins, N.F., Cadavid-Restrepo, G., de
Souza-Ault M.R., Augustyns, K., Teixeira, A.R., Schrével, J., Maigret, B., da
Silveira, J.F. & Santana JM., 2005. Molecular, functional and structural
properties of the prolyl oligopeptidase of Trypanosoma cruzi (POP Tc80) that is
required for parasite entry into mammalian cells. Biochem. J., 388: 2938.
Bennett, D.C., Cooper, P.J. & Hart, I.R., 1987. A line of non-tumorigenic
mouse melanocytes, syngeneic with the B16 melanoma and requiring a tumor
promoter for growth. Int. J. Cancer, 39: 414-418.
Beyrer, C., Villar,J.C., Suwanvanichkij, V., Singh, S., Baral, S. &
Mills, E., 2007. Neglected diseases, civil conflicts, and the right to health.
Lancet, 370: 619-627.
Boldogh, I.R. & Pon, L.A., 2006. Interactions of mitochondria with the
actin cytoskeleton. Biochim. Biophys. Acta., 1763: 450-462.
Borges, M., Cordeiro-da-Silva, A., Vergnes, B., Sereno, D. & Ouaissi,
A., 2005. Conversion of Trypanosoma cruzi Tc52 released factor to a protein
inducing apoptosis. Tissue Cell., 37: 469-478.
Bourdet-Sicard, R., Egile, C., Sansonetti, P.J. & Tran Van Nhieu, G.,
2000. Diversion of cytoskeletal processes by Shigella during invasion of
epithelial cells. Microbes Infect., 2: 813-819.
Boutayeb, A., 2007. Developing countries and neglected diseases:
challenges and perspectives. Int. J. Equity Health, 6: 1-4.
62
Brener, Z. & Chiari, E., 1963. Variações morfológicas observadas em
diferentes amostras de Trypanosoma cruzi. Rev. Inst. Med. Trop. São Paulo, 5:
220-224.
Brener Z., Costa, C.A.G. & Chiari, C., 1976. Differences in the
susceptibility of Trypanosoma cruzi strains to active chemotherapeutic agents.
Rev. Inst. Med. Trop. o Paulo, 18: 450-455.
Briones, M.R.S., Souto, R.P., Stolf, B.S.& Zingales, B., 1999. The
evolution of two Trypanosoma cruzi subgroups inferred from rRNA genes can
be correlated with the interchange of American mammalian faunas in the
Cenozoic and has implications to pathogenicity and host specificity. Mol.
Biochem. Parasitol., 104: 219-232.
Brisse, S., Dujardin, J.C. & Tibayrenc, M., 2000. Identification of six
Trypanosoma cruzi lineages by sequence-characterized amplified region
markers. Mol. Biochem. Parasitol., 111: 95-105.
Buck, A.A., Anderson, R.I. & McRae, A.A., 1978. Epidemiology of poly-
parasitism. II. Types of combinations, relative frequency and associations of
multiple infections. Tropenmed. Parasitol., 29: 137-144.
Buekens, P., Almendares, O., Carlier, Y., Dumonteil, E., Eberhard,
M., Gamboa-Leon, R., James, M., Padilla, N., Wesson, D. & Xiong, X., 2008.
Mother-to-child transmission of Chagas’ disease in North America: why dont we
do more? Matern. Child Health J., 12: 283-286.
Burleigh, B.A. & Andrews, N.W., 1998. Signaling and host cell invasion
by Trypanosoma cruzi. Curr. Opin. Microbiol., 1: 461-465.
Buscaglia, C.A., Campo, V.A., Frasch, A.C. & Di Noia, J.M., 2006.
Trypanosoma cruzi surface mucins: host-dependent coat diversity. Nat. Ver.
Microbiol., 4: 229-236.
Cabral, H.R., 2000. The tumoricidal effect of Trypanosoma cruzi: its
intracellular cycle and the immune response of the host. Med. Hypothesis, 54:
1-6.
Cardoso de Almeida, M.L. & Heise, N., 1993. Proteins anchored via
glycosylphosphatidylinositol and solubilizing phospholipases in Trypanosoma
cruzi. Biol. Res., 26: 285-312.
Carmo, M.S., Araya, J.E., Ramirez, M.I., Boscardin, S., Cano, M.I.,
Baida, R.P., Ruiz, R.C., Santos, M.R., Chiurillo, M.A., Ramirez, J.L.,
Yoshida, N. & Silveira, J.F., 1999. Organization and expression of a multigene
family encoding the surface glycoproteins of Trypanosoma cruzi metacyclic
63
trypomastigotes involved in the cell invasion. Mem. Inst. Oswaldo Cruz, 1: 169-
171.
Cerqueira, G.C, Bartholomeu, D.C., DaRocha, W.D., Hou, L., Freitas-
Silva, D.M., Machado, C.R., El-Sayed N.M. & Teixeira, S.M., 2008. Sequence
diversity and evolution of multigene families in Trypanosoma cruzi. Mol.
Biochem. Parasitol., 157: 65-72.
Chiari, E., 1981. Diferenciação do Trypanosoma cruzi em cultura. Ph.D.
Thesis. Universidade Federal de Minas Gerais, M.G., Brasil.
Chuenkova, M.V., Furnari, F.B., Cavenee, W.K. & Pereira, M.A., 2001.
Trypanosoma cruzi trans-sialidase: a potent and specific survival factor for
human Schwann cells by means of phosphatidylinositol 3-kinase/Akt signaling.
Proc. Natl. Acad. Sci. USA, 98: 9936-9941.
Cortez, M, Neira, I., Ferreira, D., Luquetti, A.O., Rassi, A., Atayde,
V.D. & Yoshida,N., 2003. Infection by Trypanosoma cruzi metacyclic forms
deficient in gp82 but expressin a related surface molecule, gp30. Infect.
Immun., 71: 6184-6191.
Cortez, M., Silva, M.R., Neira, I., Ferreira, D., Sasso, G.R., Luquetti,
A.O., Rassi, A. & Yoshida, N., 2006a. Trypanosoma cruzi surface molecule
gp90 downregulates invasion of gastric mucosal epithelium in orally infected
mice. Microbes Infect., 8: 36-44.
Cortez, M., Atayde, V.D. & Yoshida, N., 2006b. Host cell invasion
mediated by Trypanosoma cruzi surface molecule gp82 is associated with F-
actin disassembly and is inhibited by enteroinvasive Escherichia coli. Microbes
Inf., 8: 1502-1512.
Cossart, P. & Sansonetti, P.J., 2004. Bacterial invasion: the paradigms
of enteroinvasiva pathogens. Science, 304: 242-248.
Coura, J.R., Junqueira, A.C.V., Fernandes, O., Valente S.A.S. &
Miles, M.A., 2002. Emerging Chagas disease in Amazonian Brazil. Trends
Parasitol., 18: 171-176.
Coura, J.R., 2006. Transmissão da infecção chagásica por via oral na
história natural da doença de Chagas. Rev. Soc. Bras. Med. Trop., 39: 113-117.
Coura, J.R., 2007. Chagas disease: what is known and what is needed
A background article. Mem. Inst. Oswaldo Cruz, 102: 113-122.
Coutinho, M., Freire Jr., O. & Dias, J.C.P., 1999. The Noble enigma:
Chagas’ nominations for the Nobel prize. Mem. Inst. Oswaldo Cruz, 94: 123-
129.
64
Covarrubias, C., Cortez, M. Ferreira, D. & Yoshida, N., 2007.
Interaction with host factors exacerbates Trypanosoma cruzi cell invasion
capacity upon oral infection. Int. J. Parasitol., 37: 1609-1616.
Correa, M., Machado, J. Jr., Carneiro C.R., Pesquero, J.B., Bader, M.,
Travassos, L.R., Chammas, R. & Jasiulionis, M.G., 2005. Transient
inflammatory response induced by apoptotic cells is an important mediator of
melanoma cell engraftment and growth. Int. J. Cancer, 114: 356-363.
Crennell, S.J., Garman, E.F., Laver, W.G., Vimr, E.R. & Taylor, G.L.,
1993. Crystal structure of a bacterial sialidase (from Salmonella typhimurium
LT2) shows the same fold as an influenza virus neuraminidase. Proc. Natl.
Acad. Sci. USA, 90: 9852-9856.
Crisante, G., Rojas., A., Teixeira, M.G. & Añez, N., 2006. Infected dogs
as a risk factor in the transmission of human Trypanosoma cruzi infection in
western Venezuela. Acta Tropica, 98: 247-254.
Cross, G.A., & Takle, G.B., 1993. The surface trans-sialidase family of
Trypanosoma cruzi. Annu. Rev. Microbiol., 47: 385-411.
Deane, M.P., Lenzi, H.L. & Jansen, A.M., 1984. Trypanosoma cruzi:
vertebrate and invertebrate cycles in the same mammal host. Mem Inst.
Oswaldo Cruz, 79: 513-515.
Devera, R., Fernandes, O. & Coura, J.R., 2003. Should Trypanosoma
cruzi de called “cruzi complex? A review of the parasite diversity and the
potential of selecting population after in vitro culturing and mice infection. Mem.
Inst. Oswaldo Cruz, 98: 1-12.
Di Noia, J.M., D’Orso, I., Aslund, L., Sánchez, D.O. & Frasch, A.C.,
1998. The Trypanosoma cruzi mucin family is transcribed from hundreds of
genes having hypervariable regions. J. Biol. Chem., 273: 10843-10850.
Di Noia, J.M., Buscaglia, C.A., De Marchi, C.R., Almeida, I.C. &
Frasch, A.C.C., 2002. A Trypanosoma cruzi small surface molecule provides
the first immunological evidence that Chagas’ disease is due to a single
parasite lineage. J. Exp.Med., 195: 401-413.
Dias, W.B., Fajardo, F.B., Graça-Souza, A.V., Freire-de-Lima, L.,
Vieira, F., Girard, M.F., Bouteille, B., Previato, J.O., Mendonça-Previato, L.
& Todeschini, A.R., 2008. Endothelial cell signaling induced by trans-sialidase
from Trypanosoma cruzi. Cell Microbiol., 10: 88-99.
65
Docampo, R., Scott, D.A., Vercesi, A.E. & Moreno, S.N.J., 1995.
Intracellular Ca
2+
storage in acidocalcisomes of Trypanosoma cruzi. Biochem.
J,. 310: 1005-1012.
Dorta, M.L., Ferreira, A.T., Oshiro, M.E.M. and Yoshida, N., 1995.
Ca
2+
signal induced by Trypanosoma cruzi metacyclic trypomastigote surface
molecules implicated in mammalian cell invasion. Mol. Biochem. Parasitol., 73:
285-289.
Dvorak, J.A., Hall, T.E., Crane, M.S, Engel, J.C., Mc Daniel, J.P. &
Uriegas, R., 1982. Trypanosoma cruzi: flow cytometric analysis. I. Analysis of
total DNA/organism by means of mithramycin-induced fluorescence. J.
Protozool., 29: 430-437.
Elias, M.C.Q.B., Vargas, N., Tomazi, L., Pedroso, A., Zingales, B.,
Schenkman, S. & Briones, M.R.S., 2005. Comparative analysis of genomic
sequences suggests that Trypanosoma cruzi CL Brener contains two sets of
non-intercalated repeats of satellite DNA that correspond to T. cruzi I and T.
cruzi II types. Mol. Biochem. Parasitol., 140: 221-227.
El-Sayed, N.M., Myler, P.J., Bartholomeu, D.C., Nilsson, D.,
Aggarwal, G., Tran, A.N., Ghedin, E., Worthey, E.A., Delcher, A.L., Blandin,
G., Westenberger, S.J., Caler, E., Cerqueira, G.C., Branche, C., Haas, B.,
Anupama, A., Arner, E., Aslund, L., Attipoe, P., Bontempi, E., Bringaud, F.,
Burton, P., Cadag, E., Campbell, D.A., Carrington, M., Crabtree, J., Darban,
H., da Silveira, J.F., de Jong, P., Edwards, K., Englund, P.T., Fazelina, G.,
Feldblyum, T., Ferella, M., Frasch, A.C., Gull, K., Horn, D., Hou, L., Huang,
Y., Kindlund, E., Klingbeil, M., Kluge, S., Koo, H., Lacerda, D., Levin, M.J.,
Lorenzi, H., Louie, T., Machado, C.R., McCulloch, R., McKenna, A., Mizuno,
Y., Mottram, J.C., Nelson, S., Ochaya, S., Osoegawa, K., Pai, G., Parsons,
M., Pentony, M., Pettersson, U., Pop, M., Ramirez, J.L., Rinta, J.,
Robertson, L., Salzberg, S.L., Sanchez, D.O., Seyler, A., Sharma, R.,
Shetty, J., Simpson, A.J., Sisk, E., Tammi, M.T., Tarleton, R., Teixeira, S.,
Van Aken, S., Vogt, C., Ward, P.N., Wickstead, B., Wortman, J., White, O.,
Fraser, C.M., Stuart, K.D. & Andersson, B., 2005. The genome sequence of
Trypanosoma cruzi, etiologic agent of Chagas disease. Science, 309: 409-415.
Engmam, D.M., & Leon, J.S., 2002. Pathogenesis of Chagas heart
disease: role for autoimmunity. Acta. Trop., 81: 123-132.
Favoreto Jr, S., Dorta, M.L. & Yoshida, N., 1998. Trypanosoma cruzi
175-kDa protein tyrosine phosphorylation is associated with host cell invasion.
Exp. Parasitol., 89: 188-194.
66
Fecker, L.F., Geilen, C.C., Hossini, A.M., Schwarz, C., Fechner, H.,
Bartlett, D.L., Orfanos, C.E. & Eberle, J., 2005. Selective induction of
apoptosis in melanoma cells by tyrosinase promoter-controlled CD95 ligand
overexpression. J. Invest. Dermatol., 124: 221-228.
Ferreira, D., Cortez, M., Atayde, V. & Yoshida, N., 2006. Actin
cytoskeleton-dependent and independent host cell invasion by Trypanosoma
cruzi is mediated by distinct parasite surface molecules. Infect. Immun., 74:
5522-5528.
Franco, F.R., Paranhos-Bacallà, G.S., Yamauchi, L.M., Yoshida, N. &
Franco da Silveira, J., 1993. Characterization of a cDNA clone encoding the
carboxi-terminal domain of a 90-kilodalton surface antigen of Trypanosoma
cruzi metacyclic trypomastigotes. Infect. Immun., 61: 4196-4201.
Frasch, A.C., 2000. Fuctional diversity in the trans-sialidase and mucin
families in Trypanosoma cruzi. Parasitol. 16: 282-286.
Freire-de-Lima, C.G., Nunes, M.P., Corte-Real, S., Soares, M.P.,
Previato, J.O., Mendonça-Previato, L. & Dos Reis, G.A., 1998. Proapoptotic
activity of a Trypanosoma cruzi ceramide-containing glycolipid turned on in host
macrophages by IFN-gamma. J. immunol., 161: 4909-4916.
Fulda, S. & Debatin, K.M., 2006. Modulation of apoptosis signaling for
cancer therapy. Arch. Immunol. Ther. Exp., 54: 173-175.
Furstenberger, G., Barry D.L., Sorg, B. & Marks, F., 1981. Skin tumor
promotion by phorbol esters is a two-stage process. Proc. Natl. Acad. Sci., 78:
7722-7726.
Giordano, R., Fouts, D.L., Tewari, D., Colli, W., Manning, J.E., Alves,
M.J.M., 1999. Cloning of a surface membrane glycoprotein specific for the
infective form of Trypanosoma cruzi having adhesive properties to laminin. J.
Biol. Chem., 274: 3461-3468.
Gourlay, C.W. & Ayscough, K.R., 2005. The actin cytoskeleton: a key
regulator of apoptosis and ageing? Nat. Rev. Mol. Cell Biol., 6: 583-589.
Grellier, P., Vendeville, S., Joyeau, R., Bastos, I.M., Drobecq, H.,
Frappier, F., Teixeira, A.R., Schrével, J., Davioud-Charvet, E., Sergheraert,
C. & Santana, J.M., 2001. Trypanosoma cruzi prolyl oligopeptidase Tc80 is
involved in nonphagocytic mammalian cell invasion by trypomastigotes. J. Biol.
Chem., 276: 4707847080.
Hanahan, D. & Weinberg, R.A., 2000. The hallmarks of cancer. Cell,
100: 57-70.
67
Herlyn, M., Padarathsingh, M., Chin, L., Hendrix, M., Becker, D.,
Nelson, M., DeClerck, Y., McCarthy, J. & Mohla, S., 2002. New approaches
to the biology of melanoma: a workshop of the National Institutes of Health
Pathology B Study Section. Am. J. Pathol., 161: 1949-1957.
Heussler, V.T., Kuenzi, P. & Rottenberg, S., 2001. Inhibition of
apoptosis by intracellular protozoan parasites. Int. J. Parasitol., 31: 1166-1176.
Hitomi, J., Katayama, T., Taniguchi, M., Honda, A., Imaizumi, K. &
Tohyama, M., 2003. Apoptosis induced by endoplasmic reticulum stress
depends on activation of caspase-3 via caspase-12. Neurosci. Lett., 357: 127-
130.
Ho, L.H., Read, S.H., Dorstyn, L., Lambrusco, L. & Kumar, S., 2008.
Caspase-2 is required for cell death induced by cytoskeletal disruption.
Oncogene, 27: 3393-3404.
Hoft, D.F., 1996. Differential mucosal infectivity of different life stages of
Trypanosoma cruzi. Am. J. Trop. Med. Hyg., 55: 360-364.
Hoft, D.F., Farrar, P.L. & Kratz-Owens, K., D., 1996. Gastric invasion
by Trypanosoma cruzi and induction of protective mucosal immune responses.
Infect. Immun., 64: 3800-3810.
Hunter, C.A., Yu, D., Gee, M., Ngo, C.V., Sevignani, C., Goldschmidt,
M., Golovkina, T.V., Evans, S., Lee, W.F. & Thomas-Tikhonenko, A., 2001.
Cutting edge: systemic inhibition of angiogenesis underlies resistance to tumors
during acute toxoplasmosis. J. Immunol., 166: 5878-5881.
Jansen, A.M, Madeira, F., Carreira, J.C., Medina-Acosta, E. & Deane,
M.P., 1997. Trypanosoma cruzi in the opossum Didelphis marsupialis: a study
of the correlations and kinetics of the systemic and scent gland infections in
naturally and experimentally infected animals. Exp. Parasitol., 86: 37-44.
Kallinikova, V.D., Matekin, P.V., Ogloblina, T.A., Leikina, M.I.,
Kononenko, A.F., Sokolova, N.M. & Pogodina, L.S., 2001. Anticancer
properties of flagellate protozoan Trypanosoma cruzi. Izv. Akad. Nauk. Ser.
Biol., 3: 299-311.
Lages-Silva, E., Ramírez, L.E., Pedrosa, A.L., Crema, E., Galvão,
L.M.C., Pena, S.D.J., Macedo, A.M. & Chiari, E., 2006. Variability of
kinetoplast DNA gene signatures of Trypanosoma cruzi II strains from patients
with different clinical forms of Chagas’ disease in Brazil. J. Clin. Microbiol., 44:
2167-2171.
68
Leadsham, J.E. & Gourlay, C.W., 2008. Cytoskeletal induced apoptosis
in yeast. Biochim. Biophy. Acta., Epub ahead of print.
Lima, M.T., Jansen, A.M., Rondinelli, E. & Gattass, C.R., 1990.
Trypanosoma cruzi: properties of a clone isolated from CL strain. Parasitol.
Res., 77: 77-81.
Liu, H., Miller, E., Van der Water, B. & Stevens, J.L., 1998.
Endoplasmic reticulum stress proteins block oxidant-induced Ca
2+
increases
and cell death. J Biol Chem., 273: 12858-12862.
Magdesian, M., Giordano, R., Ulrich, H., Juliano, M.A., Schumacher,
R.I., Colli, W. & Alves, M.J.M., 2001. Infection by Trypanosoma cruzi:
identification of a parasite ligand and its host cell receptor. J. Biol. Chem., 276:
1938219389.
Magdesian, M., Tonelli, R., Fessel, M., Silveira, M., Schumacher, R.I.,
Linden, R., Colli, W. & Alves, M.J.M., 2007. A conserved domain of the
gp85/trans-sialidase family activates host cell extracellular signal-regulated
kinase and facilitates Trypanosoma cruzi infection. Exp. Cell Res., 313: 210-
218.
Málaga, S. & Yoshida, N., 2001. Targeted reduction in expression of
Trypanosoma cruzi surface glycoprotein gp90 increases parasite infectivity.
Infect. Immun., 69: 353-359.
Manque, P.M., Eichinger, D., Juliano, M.A., Juliano, L., Araya, J.E. &
Yoshida, N., 2000. Characterization of the cell adhesion site of Trypanosoma
cruzi metacyclic stage surface glycoprotein gp82. Infect. Immun., 68: 478-484.
Manque, P.M., Neira, I., Atayde, V.D., Cordero, E., Ferreira, A.T., da
Silveira, J.F., Ramirez, M. & Yohida, N., 2003. Cell adhesion and Ca signaling
activity in stably transfected Trypanosoma cruzi epimastigotes expressing the
metacyclic stage-specific surface molecule gp82. Infect. Immun., 71: 1561-
1565.
Mathieu-Daudé, F., Lafay, B., Touzet, O., Lelièvre, J., Parrado, F.,
Bosseno, M.F., Rojas, A.M., Fatha, S., Ouaissi, A. & Brenière, S.F., 2008.
Exploring the FL-160-CRP gene family through sequence variability of the
complement regulatory protein (CRP) expressed by the trypomastigote stage of
Trypanosoma cruzi. Infect. Genet. Evol., 8: 258-266.
Meirelles, M.N., Juliano, L., Carmona, E., Silva, S.G., Costa, E.M.,
Murta, A.C. & Scharfstein, J., 1992. Inhibitors of the major cysteinyl proteinase
(gp57/51) impair host cell invasion and arrest the intracellular development of
Trypanosoma cruzi in vitro. Mol. Biochem. Parasitol., 52: 175-184.
69
Mendonça, M.B., Nehme, N.S., Santos, S.S., Cupolillo, E., Vargas, N.,
Junqueira, A., Naiff, R.D., Barret, T.V., Coura, J.R., Zingales, B. &
Fernandes, O., 2002. Two main clusters within Trypanosoma cruzi zymodeme
3 are defined by distinct regions of the ribossomal RNA cistron. Parasitoloy,
124: 177-184.
Miles, M.A., Lanham, S.M., de Souza, A.A. & Povoa, M., 1980. Further
enzymic characters of Trypanosoma cruzi and their evaluation for strain
identification. Trans. R. Soc. Trop. Med. Hyg., 74: 221-237.
Miles, M.A., Feliciangeli., M.D. & de Arias, A.R., 2003. American
trypanosomiasis (Chagas’ disease) and the role of molecular epidemiology in
guiding control strategies. British Med. J., 28: 1444-1448.
Morel, C.M. & Lazdins, J., 2003. Chagas’ disease. Nature Rev.
Microbiol., 1: 14-15.
Moreno, S.N.J., Silva, J., Vercesi, A.E. & Docampo, R., 1994.
Cytosolic-free calcium elevation in Trypanosoma cruzi is required for cell
invasion. J. Exp. Med., 180: 1535-1540.
Mortara, R.A., Silva, S., Araguth, M.F., Blanco, S.A. & Yoshida, N.,
1992. Polymorphism of the 35- and 50-kilodalton surface glycoconjugates of
Trypanosoma cruzi metacyclic trypomastigotes. Infect. Immun., 60: 4673-4678.
Mortara, R.A., Andreoli, W.K., Taniwaki, N.N., Fernandes, A.B., Silva,
C.V., Fernandes, M.C., L’Abbate, C. & Silva, S., 2005. Mammalian cell
invasion and intracellular traficking by Trypanosoma cruzi infective forms. An.
Acad. Bras. Ciênc., 77: 77-94.
Mucci, J., Risso, M.G., Leguizamon, M.S., Frasch, A.C. &
Campetella. O., 2006. The trans-sialidase from Trypanosoma cruzi triggers
apoptosis by target cell sialylation. Cell Microbiol., 8: 1086-1095.
Murta, A.C., Persechini, P.M., Padron, T. de S., de Souza, W.,
Guimarães & J.A., Scharfstein, J., 1990. Structural and fucntional
identification of GP57/51 antigen of Trypanosoma cruzi as a cysteine
proteinase. Mol. Biochem. Parsitol., 43: 27-38.
Nde, P.N., Simmons, K.J., Kleshchenko, Y.Y., Pratap, S., Lima, M.F.
& Villalta, F., 2006. Silencing of the laminin gamma-1 gene blocks
Trypanosoma cruzi infection. Infect. Immun., 74: 1643-1648.
Neira, I., Ferreira, A.T. & Yoshida, N., 2002. Activation of distinct signal
transduction pathways in Trypanosoma cruzi isolates with differential capacity
to invade host cells. Int. J. Parasitol., 32: 405-414.
70
Neira, I., Silva, F.A., Cortez, M. & Yoshida, N., 2003. Involvement of
Trypanosoma cruzi metacyclic trypomastigote surface molecule gp82 in
adhesion to gastric mucin and invasion of epithelial cells. Infect. Immun., 71:
557-561.
Oba-Shinjo, S.M., Correa, M., Ricca, T.I., Molognoni, F., Pinhal, M.A.,
Neves, I.A., Marie, S.K., Sampaio, L.O., Nader, H.B., Chammas R. &
Jasiulionis, M.G., 2006. Melanocyte transformation associated with substrate
adhesion impediment. Neoplasia, 8: 231-241.
Odaka, C., Sanders, M.L. & Crews, P., 2000. Jasplakinolide induces
apoptosis in various transformed cell lines by a caspase-3-like protease-
dependent pathway. Clin. Diagn. Lab. Immunol., 7: 947-952.
Oliveira, E.C., Leite, M.S., Miranda J.A., Andrade, A.L., Garcia, S.B.,
Luquetti, A.O. & Moreira, H., 2001. Chronic Trypanosoma cruzi infection
associated with low incidence of 1,2-dimethylhydrazine-induced colon cancer in
rats. Carcinogenesis, 22: 737-740.
Oliveira F.O. Jr., Alves, C.R., Calvet, C.M., Toma, L., Bouças, R.I.,
Nader, H.B., Castro Côrtes, L.M., Krieger, M.A., Meirelles, M. de N. & Souza
Pereira, M.C., 2008. Trypanosoma cruzi heparin-binding proteins and the
nature of the host cell heparan sulfate-binding domain. Microb. Pathog., 44:
329-338.
Otega-Barria, E. & Pereira, M.E.A., 1991. A novel Trypanosoma cruzi
heparin binding protein promotes fibroblast adhesion and penetration on
engineered bacteria and trypanosomes into mammalian cells. Cell, 67: 411
421.
Osuna, A., Rodriguez-Cabezas, N., Boy, M., Castanys, S. & Gamarro,
F., 1993. The invasion mechanism of the metacyclic forms of Trypanosoma
cruzi in nonphagocytic host cells. Biol. Res., 26: 19-26.
Ouaissi, M.A., Afchain, D., Capron, A. & Grimaud, J.A., 1984.
Fibronectin receptors on Trypanosoma cruzi trypomastigotes and their
biological function. Nature, 308: 380-382.
Paiva, C.N., Souto-Padron, T., Costa, D.A. & Gattass, C.R., 1998.
High expression of a functional cruzipain by a non-infective and non-pathogenic
Trypanosoma cruzi clone. Parasitol., 117: 483-490.
Paiva, C.N., Castelo-Branco, M.T., Lannes-Vieira, J. & Gattass, C.R.,
1999. Trypanosoma cruzi: protective response of vaccinated mice is mediated
by CD8+ cells, prevents signs of polyclonal T lymphocyte activation, and allows
restoration of a resting immune state after challenge. Exp. Parasitol., 91: 7-19.
71
Parsons, M., Worthey, E.A., Ward, P.N. & Mottram, J.C., 2005.
Comparative analysis of the kinomes of three pathogenic trypanosomatids:
Leishmania major, Trypanosoma brucei and Trypanosoma cruzi. BMC
Genomics, 6: 127.
Pawelek, J.M., Low, K.B. & Bermudes, B., 1997. Tumor-targeted
Salmonella as a novel anticancer vector. Cancer Res., 57: 4537-4544.
Pedroso, A., Cupolillo, E. & Zingales, B., 2007. Trypanosoma cruzi:
Exploring the nuclear genome of zymodeme 3 stocks by chromosome size
polymorphism. Exp. Parasitol., 116: 71-76.
Petney, T.N. & Andrews, R.H., 1998. Multiparasite communities in
animals and humans: frequency, structure and pathogenic significance. Int. J.
Parasitol., 28: 377-393.
Posey, S.C. & Bierer, B.E., 1999. Actin stabilization by jasplakinolide
enhances apoptosis induced by cytokine deprivation. J. Biol. Chem., 274: 4259-
4265.
Previato, J.O., Jones, C., Gonçalves, L.P., Wait, R., Travassos, L.R.
& Mendonça-Previato, L., 1994. O-glycosidically linked N-acetylglucosamine-
bound oligosaccharides from glycoproteins of Trypanosoma cruzi., Biochem. J.,
301: 151-159.
Promed-mail, International Society for Infectious Diseases, 2007.
Trypanosomiasis and Chagas Disease. www.promedmail.org
Punukollu, G., Gowda, R.M., Khan, I.A., Navarro, V.S. & Vasavada,
B.C., 2007. Clinical aspects of the Chagas’ heart disease. Int. J. Cardiol., 115:
279-283.
Ramirez, M.I., Ruiz, R.C., Araya, J.E., Franco da Silveira, J. &
Yoshida, N., 1993. Involvement of the stage-specific 82-kilodalton adhesion
molecule of Trypanosoma cruzi metacyclic trypomastigotes in host cell invasion.
Infect. Immun., 61: 3636-3641.
Rassi, A., Rassi Jr., A. & Rassi, G.G., 2000. Fase Aguda. In:
Trypanosoma cruzi e doença de Chagas. Ed. Brener, Z., Andrade, Z.A., Barral-
Neto, B. Guanabara-Koogan, Rio de Janeiro.
Ratha, J., Majumdar, K.N., Mandal, S.K., Bera, R., Sarkar, C., Saha,
B., Mandal, C., Saha, K.D. & Bhadra, R., 2006. A sphingolipid rich lipid fraction
isolated from attenuated Leishmania donovani promastigote induces apoptosis
in mouse and human melanoma cells in vitro. Mol. Cell Biochem., 290: 113-123.
72
Rao, J.Y., Hurst, R.E., Bales, W.D., Jones, P.L., Bass, R.A., Archer,
L.T., Bell, B.P. & Hemstreet, G.P. 3rd., 1990. Cellular F-actin levels as a
marker for cellular transformation: relationship to cell division and differentiation.
Cancer Res., 50: 2215-2220.
Rao, V.R., Castro-Obregon, S., Frankowski, H., Schulers, M., Stoka
V., del Rio, G., Bredesen, E. & Ellerby, M.H., 2002. Coupling endoplasmic
reticulum stress to the cell death program.An Apaf-1-independent intrinsic
pathway. J Biol Chem., 277: 21836-21842.
Reddy, M., Sudeep, S.G., Kalkar, S.R., Wu, W., Anderson, P.J. &
Rochon, P.A., 2007. Oral drug therapy for multiple neglected tropical diseases.
JAMA, 298: 1911-1924.
Rey, L., 2002. Tripanossomíase devida ao Trypanosoma cruzi. In: Bases
da Parasitologia Médica. 2
a
ed. Ed. Guanabara Koogan, Cap. 3 e 4: 27-42.
Richmond, A., 2002. NF-kappa B, chemokine gene transcription and
tumour growth. Nat Rev Immunol. 2: 664-674.
Rosestolato, C.T., Dutra, J. da M., De Souza, W. & de Carvalho, T.M.,
2002. Participation of host cell actin filaments during interaction of
trypomastigote forms of Trypanosoma cruzi with host cells. Cell. Struct. Funct.,
27: 91-98.
Ruiz, R.C., Rigoni, V.L., Gonzales, J. & Yoshida, N., 1993. The 35/50
kDa antigen of Trypanosoma cruzi metacyclic trypomastigotes, an adhesion
molecule involved in host cell invasion. Parasite Immunol., 15: 121-125.
Ruiz, R.C., Favoreto Jr., S., Dorta, M.L., Oshiro, M.E.M., Ferreira,
A.T., Manque, P.M. & Yoshida, N., 1998. Infectivity of Trypanosoma cruzi
strains is associated with differential expression of surface glycoproteins with
differential Ca
2+
signaling activity. Biochem. J., 330: 505-511.
Santori, F.R., Dorta, M.L., Juliano, L., Juliano, M.A., Franco da
Silveira, J., Ruiz, R.C. & Yoshida, N., 1996. Identification of a domain of
Trypanosoma cruzi metacyclic trypomastigote surface molecule gp82 required
for attachment and invasion of mammalian cells. Mol. Biochem. Parasitol., 78:
209-21.
Satellite-Meeting, 1999. Recommendations from an International
Symposium to commemorate the 90
th
anniversary of the discovery of Chagas
disease. April 11-16 1999, Rio de Janeiro, Brazil. Mem. Inst. Oswaldo Cruz. 94:
429-432.
73
Satyamoorthy, K. & Herlyn, M., 2002. Cellular and molecular biology of
human melanoma. Cancer Biol. Ther. 1: 14-17.
Scharfstein, J., Schmitz, V., Morandi, V., Capella, M.M., Lima A.P.,
Morrot, A., Juliano, L. & Müller-Esterl, W., 2000. Host cell invasion by
Trypanosoma cruzi is potentiated by activation of bradykinin B(2) receptors. J.
Exp. Med., 192: 1289-1300.
Schenkman, S., Yoshida, N. & Cardoso de Almeida, M.L., 1988.
Glycophosphatidylinositol-anchored proteins in metacyclic trypomastigotes of
Trypanosoma cruzi. Mol. Biochem. Parasitol., 29: 141-151.
Schenkman, S., Jiang, M.S., Hart, G.W. & Nussenzweig, V., 1991a. A
novel cell surface trans-sialidase of Trypanosoma cruzi generates a stage-
specific epitope required for invasion of mammalian cells. Cell, 65: 1117-1125.
Schenkman, S., Robbins, E.S.& Nussenzweig, V., 1991b. Attachment
of Trypanosoma cruzi to mammalian cells requires parasite energy, and
invasion can be independent of the target cell cytoskeleton. Inf. Immun., 59:
645-654.
Schenkman, S., Diaz, C. & Nussenzweig, V., 1991c. Attachment of
Trypanosoma cruzi trypomastigotes to receptors at restricted cell surface
domains. Exp. Parasitol., 72: 76-86.
Schenkman, S. & Mortara, R.A., 1992. HeLa cells extend and
internalize pseudopodia during active invasion by Trypanosoma cruzi
trypomastigotes. J. Cell Sci., 101: 895-905.
Schenkman, S. & Eichinger, D., 1993. Trypanosoma cruzi trans-
sialidase and cell invasion. Parasitol. Today, 9: 218-222.
Schenkman, S., Ferguson, M.A.J., Heise, N., Cardoso de Almeida,
M.L., Mortara, R,A. & Yoshida, N., 1993. Mucin-like glycoproteins linked to the
membrane by glycosylphosphatidylinositol anchor are the major acceptors of
sialic acid in a reaction catalysed by trans-sialidase in metacyclic forms of
Trypanosoma cruzi. Mol. Biochem. Parasitol., 59: 293-304.
Schofield, C.J., 2000. Biosystematics and evolution of the Triatominae.
Cad. Saúde Pública, 16: 89-92
Schofield, C.J., Jannin, J., Salvatella., R., 2006. The future of Chagas
disease control. Trends Parasitol., 22: 583-588.
Sheklakova, L.A., Kallinikova, V.D. & Karpenko, L.P., 2003. Genetic
heterogeneity of Trypanosoma cruzi and its direct anticancer effect in cultured
human tumor cells. Bull. Exp. Biol., Med. 135: 89-92.
74
Shikanai-Yasuda, M.A., Marcondes, C.B., Guedes, L.A., Siqueira,
G.S., Barone A.A., Dias J.C., Amato Neto, V., Tolezano, J.E., Peres, B.A. &
Arruda Júnior, E.R., 1991. Possible oral transmission of acute Chagas’
disease in Brazil. Rev. Inst. Med. Trop. São Paulo., 33: 351-357.
Silva, L.H.P. & Nussenzweig, V., 1953. Sobre um cepa deTrypanosoma
cruzi altamente virulenta para o camundongo branco. F. Clin. Biol., 20: 191-
203.
Simeonova, E., Garstka, M., Koziol-Lipinska, J. & Mostowska, A.,
2004. Monitoring the mitochondrial transmembrane potential with the JC-1
fluorochrome in programmed cell death during mesophyll leaf senescence.
Protoplasma, 223: 143-153.
Songthamwat, D., Kajihara, K., Kikuchi, M., Uemura, H., Tran, S.P.M.,
Yanagi, T., Higo, H. & Hirayama, K., 2007. Structure and expression of three
gp82 gene subfamilies of Trypanosoma cruzi. Prasitol. Int. 56: 273-280.
Souto, R.P., Fernandes, O., Macedo A.M., Campbell D.A. & Zingales,
B., 1996. DNA markers define two major phylogenetic lineages of Trypanosoma
cruzi. Mol. Biochem. Parasitol. 83: 141-152.
Tardieux, I., Nathanson, M.H. & Andrews, N.W., 1994. Role in host
cell invasion of Trypanosoma cruzi-induced cytosolic free Ca
2+
transients. J.
Exp. Med., 179: 1017-1022.
Taylor, R.C., Cullen, S.P. & Martin, S.J., 2008. Apoptosis: controlled
demolition at the cellular level. Nat. Rev. Mol. Cell Biol., 9: 231-241.
Teixeira, M.G., da Silva, F.M., Marcili, A., Umezawa, E.S., Shikanai-
Yasuda, M.A., Cunha-Neto, E., Kalil, J., Stolf, N. & Stolf, A.M., 2006.
Trypanosoma cruzi lineage I in endomyocardial biopsy from a north-eastern
Brazilian patient at end-stage chronic chagasic cardiomyopathy. Trop. Med. Int.
Health, II: 294-298.
Teixeira, A.R.L., Nitz, N., Guimaro, M.C., Gomes, C. & Santos-Buch,
C.A., 2008. Chagas disease. Postgrad Med J., 82: 788-798.
Teixeira, M.M.G. & Yoshida, N., 1986. Stage-specific surface antigens
of metacyclic trypomastigotes of Trypanosoma cruzi identified by monoclonal
antibodies. Mol. Biochem. Parasitol., 18: 271-282.
Thomas, S.G., Huang, S., Li, S., Staiger, C.J. & Franklin-Tong, V.E.,
2006. Actin depolymerization is sufficient to induce programmed cell death in
self-incompatible pollen. J. Cell Biol., 174: 221-229.
75
Tsujimoto, Y., 2002. Bcl-2 family of proteins: life-or-death switch in
mitochondria. Biosci Rep., 22: 47-58.
Tzelepis, F., de Alencar, B.C., Penido, M.L., Claser, C., Machado,
A.V., Bruna-Romero, O., Gazzinelli, R.T. & Rodrigues, M.M. Infection with
Trypanosoma cruzi restricts the repertoire of parasite-specific CD8+ T cells
leading to immunodominance. J. Immunol., 180: 1737-1748.
Ueda, Y. & Richmond, A., 2006. NF-kB activation in melanoma.
Pigment Cell Res., 19: 112-124.
Ullu, E., Tschudi, C. & Gunzi, A., 1996. Trans-splicing in
trypanosomatid protozoa. In: Smith DF, Parsons M (eds). Molecular Biology of
Parasitic Protozoa. IRL Press Oxford, p. 115-129.
Van Voorhis, W.C., Barrett, L., Koelling, R. & Farr, A.G., 1993. FL-160
proteins of Trypanosoma cruzi are expressed from a multigene family and
contain two distinct epitopes that mimic nervous tissues. J. Exp. Med., 178:
681-694.
Vandenabeele, P., Orrenius, S. & Zhivotovsky, B., 2005. Serine
proteases ans calpains fulfill important supporting roles in the apoptotic tragedy
of the cellular opera. Cell Death Differ., 12: 1219-1224.
Varani, J., Wass, J.A. & Rao, K.M., 1983. Actin changes in normal
human and rat leukocytes and in transformed human leukocytic cells. J. Natl.
Cancer. Inst., 70: 805-809.
Verderame, M., Alcorta, D., Egnor, M., Smith, K. & Pollack, R., 1980.
Cytoskeletal F-actin patterns quantitated with fluorescein isothiocyanate-
phalloidin in normal and transformed cells. Proc. Natl. Acad. Sci., 77: 6624-8.
Westenberger, S.J., Barnabé, C., Campbell, D.A. & Sturn, N.R., 2005.
Two hybridization events define the population structure of Trypanosoma cruzi.
Genetics, 171: 527-543.
World Health Organization (WHO), 2005. Chagas’ Disease. Training in
Tropical Diseases (TDR): Seventeenth Program Report.
www.who.int/tdr/publications/publictions/pdf/pr17/chagas.pdf
World Health Organization (WHO), 2008. Control of Neglected Tropical
Diseases. www.who.int/neglected_diseases
Worboys, M., 1993. Tropical diseases. In: W.F. Bynum & R. Porter
(eds), Companion Encyclopedia of the History of Medicine, Routledge, London.
P. 512-536.
76
Xie, Q., Khaoustov, V.I., Chung, C.C., Sohn, J., Krishnan,B, Lewis,
D.E. & Yoffe, B., 2002. Effect of tauroursodeoxycholic acid on endoplasmic
reticulum stress-induced caspase-12 activation. Hepatology, 36: 592-601.
Yamada, T., Goto, M., Punj V., Zaborina, O., Kimbara, K., Das Gupta,
T.K. & Chakrabarty, A.M., 2002. The bacterial redox protein azurin induces
apoptosis in J774 macrophages through complex formation and stabilization of
the tumor suppressor protein p53. Infect. Immun., 70: 7054-7062.
Yamazaki, Y., Tsugura, M., Zhou, D., Fujita, Y., Shang, X., Dang, Y.,
Kawasaki, K. & Oka S., 2000. Cytoskeletal disruption accelerates caspase-3
activation and alters the intracellular membrane reorganization in DNA damage-
induced apoptosis. Exp. Cell Res., 259:64-78.
Yakubu, M.A., Majumder, S. & Kierszembaum, F., 1994. Changes in
Trypanosoma cruzi infectivity by treatments that affect calcium ion levels. Mol.
Biochem. Parasitol., 66: 119-125.
Yoshida, N., Mortara, R.A., Araguth, M.F., Gonzalez, J.C. & Russo,
M., 1989. Metacyclic neutralizing effect of monoclonal antibody 10D8 directed
to the 35 and 50 kilodalton surface glycoconjugates of Trypanosoma cruzi.
Infect. Immun., 57: 1663-1667.
Yoshida, N., Blanco, S.A., Araguth, M.F., Russo, M. & Gonzales, J.,
1990. The stage-specific 90-kilodalton surface antigen of metacyclic
trypomastigotes of Trypanosoma cruzi. Mol.Biochem. Parasitol., 39: 39-46.
Yoshida, N., Dorta, M.L., Ferreira, A.T., Oshiro, M.E.M. Mortara, R.A.,
Acosta-Serrano, A. & Favoreto Jr, S., 1997. Removal of sialic acid from
mucin-like surface molecules of Trypanosoma cruzi metacyclic trypomastigotes
enhances parasite-host cell interaction. Mol. Biochem. Parasitol., 84: 57-67.
Yoshida, N., Favoreto Jr., S., Ferreira, A.T. & Manque, PM., 2000.
Signal transduction induced in Trypanosoma cruzi metacyclic trypomastigotes
during the invasion of mammalian cells. Braz. J. Med. Biol. Research, 33: 269-
278.
Yoshida, N., 2006. Molecular basis of mammalian cell invasion by
Trypanosoma cruzi. An. Acad. Bras. Ciênc., 78: 87-111.
Yoshida, N., 2008. Trypanosoma cruzi infection by oral route: how the
interplay between parasite and host components modulates infectivity.
Parasitol. Int., 57: 105-109.
Yoshida, N. & Cortez, M., 2008. Trypanosoma cruzi: parasite and host
cell signaling during the invasion process., Subcell Biochem., 47: 82-91.
77
Zingales, B., Stolf, B.S., Souto, R.P., Fernandes, O. & Briones, M.R.,
1999. Epidemiology, biochemistry and evolution of Trypanosoma cruzi lineages
based on ribosomal RNA sequences. Mem. Inst. Oswaldo Cruz, 94: 159-164.
Molecular basis of non-virulence of Trypanosoma cruzi clone CL-14
Vanessa D. Atayde
a
, Ivan Neira
a
, Mauro Cortez
a
, Daniele Ferreira
a
,
Edna Freymu
¨
ller
b
, Nobuko Yoshida
a,
*
a
Departamento de Microbiologia, Imunologia e Parasitologia, Escola Paulista de Medicina, Universidade Federal de Sa
˜
o Paulo,
R. Botucatu, 862-6º
¯
andar, Sa
˜
o Paulo, SP 04023-062, Brazil
b
Centro de Microscopia Eletro
ˆ
nica, Escola Paulista de Medicina, Universidade Federal de Sa
˜
o Paulo,
R. Botucatu, 862-6º
¯
andar, Sa
˜
o Paulo, SP 04023-062, Brazil
Received 12 January 2004; received in revised form 5 March 2004; accepted 5 March 2004
Abstract
We investigated the properties of metacyclic trypomastigotes of non-virulent Trypanosoma cruzi clone CL-14, as compared to the parental
isolate CL. In contrast to the CL isolate, which produces high parasitemias in mice, metacyclic forms of clone CL-14 failed to produce patent
infection. In vitro, the number of clone CL-14 parasites that entered epithelial HeLa cells, after 1 h incubation, was approximately four-fold
lower than that of the CL isolate and at 72 h post-infection intracellular replication was not apparent whereas cells infected with the CL
isolate contained large number of parasites replicating as amastigotes. CL isolate metacyclic forms were long and slender, with the
kinetoplast localised closer to the nucleus than to the posterior end, whereas clone CL-14 parasites were shorter, with the kinetoplast very
close to the posterior end. Cysteine proteinase cruzipain and trans-sialidase activities were lower in CL isolate than in clone CL-14. The
surface profile was similar, except that the expression of gp82, the stage-specific glycoprotein that promotes CL isolate mucosal infection in
vivo and host cell invasion in vitro, was greatly reduced on the surface of clone CL-14 metacyclic forms. Genistein, a specific inhibitor of
protein tyrosine kinase, which is activated in CL isolate by binding of gp82 to its host cell receptor, did not affect host cell entry of clone CL-
14. In contrast with CL isolate, the infectivity of clone CL-14 was not affected by phospholipase C inhibitor U73122 but was diminished by a
combination of ionomycin plus NH
4
Cl, which releases Ca
2þ
from acidic vacuoles. Internalisation of clone CL-14, but not of CL isolate, was
significantly increased by treating parasites with neuraminidase, which removes sialic acid from the mucin-like surface molecule gp35/50.
Taken together, our data suggest an association between the non-virulence of clone CL-14 metacyclic forms and the reduced expression of
gp82, which precludes the activation of signal transduction pathways leading to effective host cell invasion.
q 2004 Australian Society for Parasitology Inc. Published by Elsevier Ltd. All rights reserved.
Keywords: Trypanosoma cruzi; Clone CL-14; Gp82; Metacyclic trypomastigotes; Cell invasion; Signal transduction
1. Introduction
The virulence of Trypanosoma cruzi, the protozoan
parasite that causes Chagas’ disease, is associated with its
ability to invade host cells and to replicate. Invasion of
mammalian cells by T. cruzi is a multi-step process that
requires the interaction of parasite and target cell molecules
and activation of signal transduction pathways in both cells.
Metacyclic trypomastigotes, the developmental forms
responsible for initiating T. cruzi infection in the mamma-
lian host, engage the stage-specific surface glycoprotein
gp82 to establish the initial parasite host cell interaction
and promote invasion (Ramirez et al., 1993; Ruiz et al.,
1998). Binding of gp82 triggers the signaling cascades in the
parasite as well as in the target cell, leading to Ca
2þ
mobilisation (Ruiz et al., 1998; Yoshida et al., 2000), which
is an essential requirement for parasite internalisation
(Moreno et al., 1994; Tardieux et al., 1994; Dorta et al.,
1995).
Experiments in mice have indicated that metacyclic
trypomastigotes have the uniquely specialised properties of
gastric mucosal invasion upon oral challenge (Hoft, 1996;
Hoft et al., 1996), a route to which is attributed the
microepidemics responsible for more than half of the acute
cases of Chagas’ disease recorded between 1968 and 2000
in Brazilian Amazon (Coura et al., 2002). Recent studies
have shown that stage-specific gp82 plays a central role in
0020-7519/$30.00 q 2004 Australian Society for Parasitology Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijpara.2004.03.003
International Journal for Parasitology 34 (2004) 851–860
www.parasitology-online.com
*
Corresponding author. Tel.: þ55-11-5576-4532; fax: þ 55-11-5571-
1095.
E-mail address: [email protected] (N. Yoshida).
mucosal invasion of metacyclic forms, leading to systemic
infection upon oral inoculation (Neira et al., 2003).
T. cruzi isolate CL, which is highly infective in vitro and
in vivo (Yoshida, 1983; Ruiz et al., 1998; Neira et al., 2003)
has been used in experiments to establish the role of gp82 in
host cell invasion (Ruiz et al., 1998; Favoreto et al., 1998;
Yoshida et al., 2000). What is intriguing is that CL-14, a
clone derived from the CL isolate, is unable to produce a
patent infection even when injected into new-born mice
(Lima et al., 1990), negative parasitism resulting from
extensive histopathological analysis of mouse tissues and
organs, at different times after intraperitoneal or intravenous
injection of metacyclic forms (Lima et al., 1995). As
different T. cruzi isolates may differentially engage their
repertoire of surface molecules to interact with host cells,
activating distinct signal transduction pathways that deter-
mine the efficiency of parasite internalisation (Ruiz et al.,
1998; Neira et al., 2002), the deficient infectivity of clone
CL-14 could be associated with its surface profile.
Investigating the molecular basis of lack of virulence of
clone CL-14 may be important to further clarify the
complex process of T. cruzi infection. To this end, we
analysed in this study the properties of metacyclic forms of
clone CL-14, as compared to those of CL isolate, focusing
on the profile of surface molecules, gp82 in particular; the
activities of cruzipain and trans-sialidase, enzymes that
have been implicated in target cell penetration (Meirelles
et al., 1992; Schenkman et al., 1991); and the ability to
invade epithelial HeLa cells upon treatment of parasites
with drugs that interfere with signal transduction and
intracellular Ca
2þ
mobilisation.
2. Materials and methods
2.1. Trypanosoma cruzi, mammalian cells
and cell invasion assay
The T. cruzi isolate CL (Brener and Chiari, 1963) was
maintained cyclically in mice and in axenic culture in liver
infusion tryptose (LIT) medium and the clone CL-14
(Chiari, 1981) was grown in LIT medium throughout this
study. To obtain cultures enriched in metacyclic trypomas-
tigotes, Grace’s medium (Life Technologies) was also used.
Metacyclic forms from cultures at the stationary growth
phase were purified by passage through DEAE-cellulose
column, as described (Teixeira and Yoshida, 1986). HeLa
cells, the human carcinoma-derived epithelial cells, were
grown at 37 8C in Dulbecco’s minimum essential medium
(DMEM) supplemented with 10% fetal calf serum,
streptomycin (100 mg/ml) and penicillin (100 U/ml) in a
humidified 5% CO
2
atmosphere. Host cell invasion assays
were carried out as detailed elsewhere (Yoshida et al.,
1989), by seeding the parasites onto each well of 24-well
plates containing 13-mm diameter round glass coverslips
coated with 1.5 £ 10
5
HeLa cells. After varying periods of
time, depending on the experiment, the duplicate coverslips
were washed in PBS and stained with Giemsa. The number
of intracellular parasites was counted in at least 500 Giemsa
stained cells.
2.2. Oral infection of mice with T. cruzi
Four-week-old female Balb/c mice were inoculated
orally with purified metacyclic trypomastigotes. The
parasites were introduced by intrapharyngeal route through
a plastic tube adapted to a 1 ml plastic syringe. Starting on
day 13 p.i., parasitemia was monitored by examining 5 ml
peripheral blood samples under phase contrast microscope,
twice a week.
2.3. Flow cytometry
Live metacyclic trypomastigotes (4 £ 10
7
) were incu-
bated for 1 h on ice, with monoclonal antibodies (MAbs)
directed to different T. cruzi surface molecules, or with
unrelated MAb 1C3 directed to Leishmania amazonensis
gp63 (Barbie
´
ri et al., 1993). After washings in PBS, the
parasites were fixed with 2% paraformaldehyde in PBS for
30 min. The fixative was washed out, and the parasites were
incubated with fluorescein-labeled goat anti-mouse IgG for
1 h at room temperature. Following two more washes, the
number of fluorescent parasites was estimated with a Becton
Dickinson FACscan cytometer. Assays with fixed and
permeabilised parasites were carried out as follows: fixation
with 2% paraformaldehyde, washings in PBS, treatment
with 0.1% saponin in PBS at room temperature for 30 min,
washings in PBS, incubation with MAb 3F6 for 1 h at room
temperature, washes in PBS, and incubation with fluor-
escein-conjugated antibody as described above.
2.4. TEM and immunocytochemistry
Purified metacyclic trypomastigotes (5 £ 10
7
)were
washed in PBS and fixed with a solution containing 0.1%
glutaraldehyde and 4% paraformaldehyde in 0.1 M sodium
cacodylate buffer, pH 7.2. After 1 h at room temperature and
washings in cacodylate buffer, the parasites were treated
with PBS containing 0.1 M glycine, for 30 min at room
temperature, to block free aldehyde groups and then washed
in PBS. The parasites were blocked with PBS containing 5%
BSA for 1 h at room temperature and incubated with MAb
3F6 or with unrelated antibody, diluted 1:10 in blocking
solution, for 2 h at room temperature. Following washings
in PBS containing 1% BSA, the parasites were incubated for
2 h with anti-mouse IgG coupled to 10 nm colloidal gold
particles diluted 1:40 in blocking solution. After inclusion in
2% agarose and treatment with 1% OsO
4
for 1 h,
dehydration in ethanol series, the parasites were embedded
in Epon 812 (EMS). Thin sections were collected on
Formvar/carbon-coated copper grids, stained with 2%
uranyl acetate for 8 min and 1% lead citrate for 4 min,
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860852
washed, dried and observed in transmission electron
microscope (Jeol 1200 EX II).
2.5. Detection of proteinase and trans-sialidase activities
T. cruzi proteinase activity was determined as follows:
3 £ 10
7
metacyclic trypomastigotes were lysed in 0.4%
Triton X-100, solubilised in sample buffer without 2-
mercaptoethanol, and then loaded onto 10% SDS-poly-
acrylamide gel containing 0.1% copolymerised gelatin as
substrate. After electrophoresis, the gels were subjected to
two 30 min washes with 2.5% Triton X-100 in 0.1 M acetate
buffer, pH 6.0, to remove SDS, and incubated overnight in
the same buffer without detergent. The gel was stained with
Coomassie blue R250 and destained for visualisation of
white bands against a blue background. Trans-sialidase
activity was determined as previously described (Schenk-
man et al., 1992). Briefly, 10
8
metacyclic forms were
pelleted, detergent lysed and processed for measurement of
trans-sialidase activity, by the transfer of sialic acid from
sialyllactose to
D-glucose-1-[
14
C]lactose and detection of
sialylated products by chromatography in QAE-Sephadex
A-25 (Pharmacia).
2.6. Treatment of parasites with neuraminidase and drugs
that interfere with cell signaling
Metacyclic trypomastigotes were treated at 37 8C for 1 h
with 0.2 U/ml of neuraminidase (type III from Vibrio
cholerae or type X from Clostridium perfringens, Sigma) in
PBS, pH 6.0, containing 1 mM CaCl
2
. After washings in
PBS, the parasites were used for reactivity towards lectin
and for cell invasion assays. Treatment of metacyclic forms
with different drugs included incubation with 250 mM
genistein at 37 8C for 30 min, 1 mM U73122 at 37 8C for
4 min; 1 mM ionomycin plus 20 mM NH
4
Cl for 10 min at
37 8C.
2.7. Southern blot analysis
T. cruzi DNA was digested with different restriction
enzymes, separated by electrophoresis on 0.8% agarose gel
and blotted onto nylon membranes. Hybridisation with the
probe, which consisted of a DNA fragment corresponding to
ORF of gp82 gene (whole insert of gp82 cDNA clone)
labeled with [
32
P], and washings were performed as detailed
(Araya et al., 1994).
2.8. Pulsed field gel electrophoresis
Agarose blocks containing genomic DNA from 10
8
parasites were prepared, incubated at 50 8C for 16 h in lysis
solution containing 10 mM Tris HCl, pH 8.0, 500 mM
EDTA, 1% sarkosyl, 1 mg/ml proteinase K, equilibrated in
TE, washed and stored in 0.5 M EDTA at 4 8C. Small
portions (equivalent to 10
7
parasites) were electrophoresed
(1.2% agarose gel in 0.5 £ TBE) at 80 V for 132 h in Gene
Navigatore System (Pharmacia), from pulse times varying
from 90 to 800 s. DNA from Hansenula wingei was used as
reference. After transfer to nylon membranes, chromosomal
DNA bands were hybridised with the [
32
P]-labeled insert of
gp82 cDNA clone and revealed by exposure to X-ray film
(Hyperfilm-MP, Amersham).
2.9. Statistics
Student’s t-test was used to determine significance in T.
cruzi cell invasion assays, in which the infectivity between
CL isolate and clone CL-14 was compared or the effect of
treatment of parasites with neuraminidase or drugs that
inhibit signal transduction was evaluated.
3. Results
3.1. In vivo and in vitro infection by metacyclic forms
of T. cruzi isolate CL and clone CL-14
We compared the infectivity of CL isolate and clone
CL-14 by oral administration, a highly efficient route
leading to systemic T. cruzi infection (Hoft, 1996; Neira
et al., 2003). Balb/c mice were injected orally with
2 £ 10
6
metacyclic forms of clone CL-14 or 4 £ 10
5
parasites of CL isolate. Starting on day 13 p.i., blood
samples were examined twice a week. Parasitemia was
not detectable in mice inoculated with clone CL-14
whereas the animals inoculated with five-fold fewer
metacyclic forms of CL isolate developed high para-
sitemias (Fig. 1A). The parasites used in these exper-
iments were . 95% pure typical metacyclic
trypomastigotes, with the kinetoplast posterior to the
nucleus, as visualised in Giemsa-stained preparations
(Fig. 2). We have noted that CL isolate metacyclic forms
were long and slender, with the kinetoplast localised
closer to the nucleus than to the posterior end, whereas
clone CL-14 parasites were shorter, with the kinetoplast
very close to the posterior end.
To examine whether the low infectivity of clone CL-
14 was associated with the impaired ability to invade
host cells, we performed invasion assays by incubating
metacyclic forms with HeLa cells for 1 h at 37 8C, at
parasite:cell ratio of 10:1. The number of internalised
parasites of clone CL-14 was approximately four-fold
lower as compared to that of the CL isolate (Fig. 1B), a
difference extremely significant ðP , 0:0001Þ: Exper-
iments were also performed to determine the ability of
clone CL-14 to replicate intracellularly. Parasites were
incubated with HeLa cells for 1 h and, after washings in
PBS, one set of duplicate coverslips were immediately
stained with Giemsa whereas other sets were reconsti-
tuted with DMEM containing 2% fetal calf serum
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860 853
and maintained for up to 72 h. No apparent intracellular
replication of clone CL-14 was observed by 72 h p.i., as
opposed to the CL isolate (Fig. 1C) that multiplied
actively as amastigotes.
3.2. Differential expression of gp82 on the surface of CL
isolate and clone CL-14
In CL isolate, gp82 is the main surface molecule of
metacyclic forms associated with the parasite infectivity
(Ramirez et al., 1993; Neira et al., 2003). We examined
whether clone CL-14 differed from CL isolate as regards
the expression of gp82. Analysis of metacyclic forms by
flow cytometry, upon reaction of live parasites with MAb
3F6, consistently revealed in CL isolate a major
population expressing high gp82 levels and a minor
population with lower fluorescence intensity, whereas in
clone CL-14 the parasites reacting poorly with MAb 3F6
predominated over those with higher gp82 expression
(Fig. 3A). We confirmed the reduced expression of gp82
Fig. 1. Differential infectivity of metacyclic forms of Trypanosoma
cruzi CL isolate and clone CL-14. (A) The course of infection in mice
inoculated with parasites by oral route was monitored by counting the
number of parasites in 5 ml blood samples. Each data point corresponds
to the mean parasitemia of five animals in each group of mice. Note the
high parasitemias produced by CL isolate in contrast to the lack of
patent infection by clone CL-14. (B) Host cell invasion was assayed by
incubating parasites with HeLa cells at 37 8C for 1 h. After washings in
PBS, the number of intracellular parasites was counted in a total of at
least 500 Giemsa-stained cells. The difference in infectivity between CL
isolate and clone CL-14 was extremely significant ðP , 0:0001Þ by
Student’s t-test. (C) Intracellular T. cruzi replication was examined by
incubating parasites with HeLa cells for 1 h. After washings in PBS,
one set of coverslips were stained with Giemsa whereas another set was
maintained for 72 h in Dulbecco’s minimum essential medium contain-
ing 2% FCS. Values are the means ^ SD of eight experiments (B) and
three experiments (C), performed in duplicate for each experiment.
Fig. 2. Metacyclic trypomastigotes of Trypanosoma cruzi CL isolate and
clone CL-14. Purified parasites were stained with Giemsa. Note the position
of kinetoplast very close to the posterior end in clone CL-14 and more
proximal to the nucleus in CL isolate.
Fig. 3. Expression of surface glycoproteins gp82 in metacyclic forms of
Trypanosoma cruzi CL isolate and clone CL-14. (A) Live parasites were
reacted with MAb 3F6 directed to gp82, fixed and then processed as
described in Section 2 for analysis by flow cytometry. Representative
results of at least seven experiments are shown. Note the lower fluorescence
of clone CL-14, reflex of reduced gp82 expression. (B) Detergent extracts,
equivalent to 3 £ 10
7
parasites, were electrophoresed in 10% SDS-PAGE
gels and analysed by immunoblotting using MAb 3F6. Gp82 bands of
comparable intensity were revealed in CL isolate and clone CL-14. (C)
Parasites were permeabilised or not by treatment with saponin before
reaction with MAb 3F6 and processing for analysis by flow cytometry.
Representative results of three experiments are shown. Note the increased
gp82 levels in permeabilised clone CL-14 parasites.
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860854
on the surface of clone CL-14 metacyclic forms by
immunocytochemical reactions visualised at the electron
microscope (Fig. 4). In CL isolate, the presence of gp82
was revealed by MAb 3F6-colloidal gold complex
throughout the parasite membrane, including the flagel-
lum, whereas in clone CL-14 metacyclic forms
the labeling was sparse. However, metacyclic forms of
CL isolate and clone CL-14 displayed gp82 bands of
comparable intensity when analysed by immunoblotting
(Fig. 3B), a method that uses extracts of detergent lysed
parasites and therefore can also detect internally localised
gp82, not accessible to MAb 3F6 in live CL-14
Fig. 4. Distribution of surface molecule gp82 in metacyclic forms of Trypanosoma cruzi CL isolate and clone CL-14, revealed by immunogold. Parasites were
incubated sequentially with MAb 3F6 directed to gp82 and anti-mouse IgG coupled to colloidal gold particles and then processed for electron microscopy. Note
the labeling of gp82 in CL isolate metacyclic forms (arrows). K, kinetoplast. Bar, 0.5 mm. Insets show the position of kinetoplast relative to the nucleus (N) and
to the posterior end of the parasite. Bar, 1 mm.
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860 855
metacyclic forms. Accordingly, permeabilisation of CL-
14 parasites with saponin before reaction with MAb 3F6
increased the gp82 levels detectable by flow cytometry,
whereas only a minor alteration in the profile of gp82
expression was observed in permeabilised CL isolate
metacyclic forms as compared to the controls (Fig. 3C).
We also compared the genomic organisation of gp82
gene family in CL isolate and clone CL-14. When
hybridised with the insert of gp82 cDNA clone, Southern
blots of genomic DNA digested with restriction enzyme
Bam HI, Eco RI, Hind III or Hae III, displayed similar
profiles in CL isolate and clone CL-14 (Fig. 5A).
Chromosomal mapping of gp82 genes, carried out by
hybridising the same probe with chromosomal size
fragments separated by pulsed field gel electrophoresis,
revealed some differences. A weak band of , 3.3 Mb and
a , 2.1 Mb fragment of high intensity were detected only
in CL isolate, whereas clone CL-14 displayed
an exclusive fragment of , 2.9 Mb and a , 1.7 Mb
band of much higher intensity than that of CL isolate
(Fig. 5B).
3.3. Expression of mucin-like surface molecule gp35/50
and of gp90 in CL isolate and clone CL-14
We also examined the profile of other metacyclic
trypomastigote surface molecules that are known to interact
with host cells. By flow cytometry, CL isolate and clone CL-
14 parasites were found to express on the surface
comparable levels of the mucin-like glycoprotein gp35/50
identified by MAb 2B10 (Fig. 6A) and immunoblots
revealed gp35/50 bands of similar intensity but displaying
some difference in size (Fig. 6B). The variant form of gp35/
50 as well as gp90, the stage-specific glycoprotein that
negatively modulates cell invasion (Ma
´
laga and Yoshida,
2001), recognised, respectively, by MAb 10D8 and MAb
1G7 in a number of poorly infective T. cruzi isolates (Ruiz
et al., 1998), were not detected in CL isolate or clone CL-14,
but both expressed a variant gp90 detectable by MAb 5E7
Fig. 5. Genomic organisation of gp82 genes in Trypanosoma cruzi CL
isolate and clone CL-14. (A) Southern blot of genomic DNA digested with
the indicated restriction enzyme was hybridised with the whole insert of
gp82 cDNA clone. Samples 1, 2, 3 and 4 correspond to CL isolate and l
0
,2
0
,
3
0
and 4
0
to clone CL-14. (B) Chromosomal bands of parasites were
separated by pulsed field gel electrophoresis, transferred to nylon
membrane and hybridised with the labeled insert of gp82 cDNA clone.
Numbers correspond to molecular sizes. Arrows point to fragments
detectable only in CL isolate and arrowheads to bands exclusive for or
that appear in higher intensity in clone CL-14.
Fig. 6. Molecular profile of metacyclic forms of Trypanosoma cruzi CL
isolate and clone CL-14. (A) Live parasites were reacted with MAb 2B10
directed to gp35/50, fixed and processed for analysis by flow cytometry.
Representative results of seven experiments are shown. The comparable
fluorescence intensity indicates that there is no difference in gp35/50
expression. (B) Detergent extracts, equivalent to 3 £ 10
7
parasites, were
electrophoresed in 10% SDS-PAGE gels and analysed by immunoblotting
with the indicated monoclonal antibodies. (C) For detection of proteases,
Triton X-100 extracts, equivalent to 3 £ 10
7
parasites, pre-incubated or not
with 100 mM E-64, were electrophoresed in 10% SDS-PAGE gel
containing gelatin. Afterwards, the gel was washed, incubated overnight
and stained with Coomassie blue. Note the high expression of cysteine
protease activity in clone CL-14.
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860856
(Fig. 6B) directed to an epitope that is cryptic in live
parasites (Teixeira and Yoshida, 1986).
3.4. Cruzipain and trans-sialidase activities in CL isolate
and clone CL-14
Cruzipain, the major T. cruzi cysteine proteinase
(gp57/51) has been implicated in host cell invasion of
tissue culture trypomastigotes as well as in the replication of
amastigotes (Meirelles et al., 1992). We determined the
cruzipain activity of CL isolate and clone CL-14 metacyclic
forms in gelatin gels, by loading parasite preparations in
absence or in the presence of 20 mM E-64, an inhibitor of
cysteine proteinase. Expression of functional cruzipain,
inhibitable by E-64, was higher in clone CL-14 (Fig. 6C). In
addition to the gp57/51 bands, a minor band migrating
slower than BSA and resistant to E-64 was detected in clone
CL-14. As regards trans-sialidase, also implicated in cell
invasion of tissue culture trypomastigotes (Schenkman et al.,
1991), the activity in clone CL-14 was , 1.6-fold higher
than in CL isolate.
3.5. Differential effect of neuraminidase treatment of CL
isolate and clone CL-14 metacyclic forms on host cell
invasion
A previous study had shown that, depending on the T.
cruzi isolate, treatment of metacyclic trypomastigotes with
neuraminidase, which removes sialic acid from mucin-like
molecule gp35/50, increases the ability to invade target cells
(Yoshida et al., 1997). To determine whether removal of
sialic acid affected the infectivity of clone CL-14 metacyclic
forms, the parasites were treated with neuraminidase as
described in Section 2. Enzyme-treated parasites had
augmented their reactivity with Bauhinia purpurea lectin,
which has affinity for
D-galactose, an indication that sialic
acid was removed, exposing galactosyl residues. As shown
in Fig. 7A, the infectivity of clone CL-14 metacyclic forms
Fig. 7. Differential effect of treatment of Trypanosoma cruzi metacyclic forms with different drugs on target cell invasion of CL isolate and clone CL-14. (A)
Parasites were either untreated or treated with neuraminidase (0.2 U/ml) for 1 h, washed and then seeded onto HeLa cells. After 1 h incubation at 37 8C,
followed by washings in PBS, the number of intracellular parasites was counted in a total of at least 500 Giemsa-stained cells. Values are the means ^ SD of
four experiments performed in duplicate. The difference between untreated controls and neuraminidase-treated CL-14 parasites was significant, whether the
neuraminidase used was from Vibrio cholerae ðP , 0:05Þ or from Clostridium perfringens ðP , 0:005Þ; by Student’s t-test. (B) Parasites, pre-treated or not
with the indicated drug, were incubated with HeLa cells, at parasite:cell ratio of 50:1 for clone CL-14 and 10:1 for CL isolate. After 30 min at 37 8C, the number
of intracellular parasites was counted in at least 500 Giemsa-stained cells. Values are the means ^ SD of three experiments performed in duplicated. For CL
isolate, a significant inhibition of cell invasion was observed upon treatment of parasites with genistein ðP , 0:0001Þ or U73122 ðP , 0:0001Þ; whereas for
clone CL-14 the difference between untreated controls and parasites treated with the combination ionomycin þ NH
4
Cl was significant ðP , 0:05Þ; by
Student’s t-test. This suggests phospholipase C-mediated Ca
2þ
release in CL isolate and phospholipase C-independent Ca
2þ
mobilisation from
acidocalcisomes in clone CL-14.
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860 857
towards HeLa cells increased significantly upon treatment
with neuraminidase from Vibrio cholerae ðP , 0:05Þ or
Clostridium perfringens ðP , 0:005Þ; whereas that of CL
isolate remained unaltered. Neuraminidase-treated clone
CL-14 parasites failed to replicate intracellularly.
3.6. Activation of distinct signaling pathways in metacyclic
forms of CL isolate and clone CL-14 during host cell
invasion
In this set of short time (30 min) cell invasion assays, in
which parasites were pre-treated with drugs that act on
signal transduction pathways leading to Ca
2þ
mobilisation,
the parasite:cell ratio used for clone CL-14 was 50:1.
Genistein, a specific inhibitor of protein tyrosine kinase
(Akiyama et al., 1987), that inhibits Ca
2þ
response and the
infectivity of CL isolate metacyclic forms (Yoshida et al.,
2000), did not affect HeLa cell entry of clone CL-14
parasites (Fig. 6B). Treatment with U73122, a specific
inhibitor of phospholipase C (Bleasdale et al., 1990), which
mediates inositol 1,4,5-triphosphate production (Berridge,
1993), inhibited HeLa cell invasion of CL isolate but not of
clone CL-14 metacyclic forms, whereas Ca
2þ
ionophore
ionomycin plus NH
4
Cl, a combination that potentiates Ca
2þ
release from acidocalcisomes (Docampo et al., 1995), had
an opposite effect, by affecting only clone CL-14 internal-
isation (Fig. 7B).
4. Discussion
Our study supports the hypothesis that the metacyclic
stage-specific surface glycoprotein gp82 is a key molecule in
promoting efficient T. cruzi infection. Of the various parasite
molecules implicated in host cell invasion analysed in this
study, gp82 was found to be expressed at much lower levels on
thesurface ofclone CL-14, ascompared tothehighly infective
parental CL isolate, but otherwise the molecular profiles of the
isolate and the clone were similar. In CL isolate, gp82 plays a
central role in target cell penetration in vitro (Ramirez et al.,
1993; Ruiz et al., 1998) as well as in promoting mucosal
infection, leading to high parasitemias, upon oral challenge
(Neira et al., 2003). The expression of reduced levels of
surface gp82 in metacyclic froms of clone CL-14 could
explain the poor cell invasive capacity of these parasites in
vitro and why they do not produce patent infection in mice
when administered orally. Gp82, which binds to gastric mucin
in a dose-dependent manner (Neira et al., 2003), may first
adhere to mucin that lines the mucosal surfaces, in order to
initiate mucosal infection. If engagement of gp82 is required
to adhere to and traverse the mucin coat, before parasite
penetration into underlying epithelial cells, the reduced
expression of gp82 in CL-14 metacyclic forms would be an
impediment for this initial interaction with the host.
Based on several pieces of evidence, we had proposed
that the most efficient target cell penetration of metacyclic
trypomastigotes is mediated by gp82, which triggers in both
cells the activation of signaling cascades leading to Ca
2þ
mobilisation from intracellular reservoirs (Ruiz et al., 1998;
Neira et al., 2003). Binding of gp82 to its receptor on the
host cell induces in CL isolate metacyclic forms the
activation of protein tyrosine kinase and Ca
2þ
release,
probably from endoplasmic reticulum, in a manner
dependent on inositol 1,4,5-triphosphate (Yoshida et al.,
2000). In accord with this, cell invasion of clone CL-14 that
is deficient in the expression of surface gp82 is not
dependent on protein tyrosine kinase. An alternative signal
transduction pathway must be induced in metacyclic forms
of clone CL-14 during target cell entry in vitro. One
possibility is that the mucin-like surface glycoprotein gp35/
50 is involved.
Gp35/50, which also induce target cell Ca
2þ
response,
although to a lesser degree than gp82 (Dorta et al., 1995),
has been implicated in host cell invasion of metacyclic
forms of poorly infective T. cruzi isolates (Ruiz et al., 1998).
This molecule is the main acceptor of sialic acid in a
reaction mediated by trans-sialidase (Schenkman et al.,
1993), but sialic acid appears to impair, rather than promote,
the interaction of gp35/50 with target cells. Treatment of
metacyclic forms with neuraminidase removes sialic acid
from gp35/50, increases the Ca
2þ
signal-inducing activity
and, depending on the T. cruzi isolate, augments the parasite
infectivity (Yoshida et al., 1997). Neuraminidase treatment
of CL-14 metacyclic forms increased the reactivity with
lectin recognising
D-galactose as well as the invasive
capacity of clone CL-14 metacyclic forms towards HeLa
cells, suggesting that the removal of sialic acid from gp35/
50 facilitates parasite target cell interaction by exposing
D-
galactose residues, which putatively are required for
recognition of target cell receptors. Neuraminidase treat-
ment did not affect CL isolate infectivity, which is not
dependent on gp35/50.
Our results indicate that distinct signaling cascades are
activated in CL isolate and clone CL-14 parasites during
host cell invasion, leading to Ca
2þ
mobilisation from
different intracellular reservoirs. Ca
2þ
release in CL
isolate, possibly from endoplasmic reticulum (Neira et al.,
2002) is mediated by inositol 1,4,5-triphosphate gener-
ated by phospholipase C, an enzyme whose inhibition
results in decreased parasite infectivity. In clone CL-14,
the Ca
2þ
necessary for parasite internalisation appears,
at least in part, to be mobilised from acidocalcisomes in
a manner independent of inositol 1,4,5-triphosphate
(Table 1).
In addition to a reduced ability to enter host cells, clone
CL-14 also appears to have impaired capacity for intra-
cellular development (Fig. 1C). As T. cruzi cysteine
proteinase cruzipain plays a role in intracellular parasite
replication (Meirelles et al., 1992), one possibility could be
a deficient replication due to the reduced cruzipain activity.
However, this is apparently not the case. Expression of
functional cruzipain was higher in clone CL-14 as compared
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860858
to CL isolate. In addition, Paiva et al. (1998) have found
similar expression of cruzipain on flagellar and cellular
membranes of trypomastigotes of CL isolate and clone CL-
14. Another possibility could be an impaired capacity to
escape from the parasitophorous vacuole to the cytoplasm,
due to a deficient expression of trans-sialidase, an enzyme
that also displays neuraminidase activity (Schenkman et al.,
1992). According to Hall et al. (1992), T. cruzi neurami-
nidase facilitates the parasite’s escape into the cytoplasm by
desialylating the vacuole membrane and rendering it more
susceptible to lysis by a parasite hemolysin. That possibility
is not compatible with trans-sialidase activity of clone CL-
14, which was , 1.6-fold higher than that of the CL isolate.
An intriguing question is whether the deficient expression in
clone CL-14 of gp82, which is a member of T. cruzi gp85/
sialidase family (Araya et al., 1994) but is apparently devoid
of enzymatic activity, impairs functions essential for
parasite development other than the ability to bind to
gastric mucin and to enter epithelial cells.
Taken together, our data suggest that the non-virulence
of clone CL-14 metacyclic forms is associated with the
reduced expression on the surface of gp82, the molecule
involved in mucosal infection in vivo and activation of
parasite signaling pathway leading to efficient host cell
invasion.
Acknowledgements
This work was supported by Fundac¸a
˜
o de Amparo a
`
Pesquisa do Estado de Sa
˜
o Paulo (FAPESP). We thank
Evania B. Azevedo for help in the detection of parasite
trans-sialidase activity, and Dr Tania A.T. Gomes for
reading the manuscript.
References
Akiyama, T., Ishida, J., Nakagawa, S., Ogawara, H., Watanabe, S., Itoh, N.,
Shibuya, M., Fukami, Y., 1987. Genistein, a specific inhibitor of
tyrosine-specific protein kinases. J. Biol. Chem. 262, 5592 5595.
Araya, J.E., Cano, M.I., Yoshida, N., Franco da Silveira, J., 1994. Cloning
and characterization of a gene for the stage-specific 82-kDa surface
antigen of metacyclic trypomastigotes of Trypanosoma cruzi. Mol.
Biochem. Parasitol. 65, 161 169.
Barbie
´
ri, C.L., Giorgio, S., Merjan, A.J., Figueiredo, E.N., 1993. Glyco-
sphingolipid antigens of Leishmania (Leishmania) amazonensis
amastigotes identified by use of a monoclonal antibody. Infect.
Immun. 61, 21312137.
Berridge, M.J., 1993. Inositol triphosphate and calcium signalling. Nature
361, 315325.
Bleasdale, J.E., Thakur, R., Gremban, R.S., Bundy, G.L., Fitzpatrick, F.A.,
Smith, R.J., Bunting, S., 1990. Selective inhibition of receptor-coupled
phospholipase C-dependent processes in human platelets and poly-
morphonuclear neutrophils. Exp. Ther. 255, 756 768.
Brener, Z., Chiari, E., 1963. Variac¸o
˜
es morfolo
´
gicas observadas em
diferentes amostras de Trypanosoma cruzi. Rev. Inst. Med. Trop. Sa
˜
o
Paulo 5, 220 224.
Chiari, E., 1981. Diferenciac¸a
˜
odoTrypanosoma cruzi em cultura. PhD
Thesis, Universidade Federal de Minas Gerais, M.G., Brasil.
Coura, J.R., Junqueira, A.C.V., Fernandes, O., Valente, S.A.S., Miles,
M.A., 2002. Emerging Chagas disease in Amazonian Brazil. Trends
Parasitol. 18, 171176.
Docampo, R., Scott, D.A., Vercesi, A.E., Moreno, S.N.J., 1995.
Intracellular Ca
2þ
storage in acidocalcisomes of Trypanosoma cruzi.
Biochem. J. 310, 1005 1012.
Dorta, M.L., Ferreira, A.T., Oshiro, M.E.M., Yoshida, N., 1995. Ca
2þ
signal induced by Trypanosoma cruzi metacyclic trypomastigote
surface molecules implicated in mammalian cell invasion. Mol.
Biochem. Parasitol. 73, 285 289.
Favoreto, S. Jr., Dorta, M.L., Yoshida, N., 1998. Trypanosoma cruzi 175-
kDa protein tyrosine phosphorylation is associated with host cell
invasion. Exp. Parasitol. 89, 188194.
Hall, B.F., Webster, P., Ma, A.K., Joiner, K.A., Andrews, N.W., 1992.
Desialylation of lysosomal glycoprotein by Trypanosoma cruzi: a role
for the surface neuraminidase in facilitating parasite entry into the host
cell cytoplasm. J. Exp. Med. 176, 313325.
Hoft, D.F., 1996. Differential mucosal infectivity of different life stages of
Trypanosoma cruzi. Am. J. Trop. Med. Hyg. 55, 360 364.
Hoft, D.F., Farrar, P.L., Kratz-Owens, K., Shaffer, D., 1996. Gastric
invasion by Trypanosoma cruzi and induction of protective mucosal
immune responses. Infect. Immun. 64, 3800 3810.
Lima, M.T., Jansen, A.M., Rondinelli, E., Gattass, C.R., 1990. Trypano-
soma cruzi: properties of a clone isolated from CL strain. Parasitol. Res.
77, 7781.
Lima, M.T., Lenzi, H.L., Gattass, C.R., 1995. Negative tissue parasitism
with a noninfective clone of Trypanosoma cruzi. Parasitol. Res. 81,
612.
Ma
´
laga, S., Yoshida, N., 2001. Targeted reduction in expression of
Trypanosoma cruzi surface glycoprotein gp90 increases parasite
infectivity. Infect. Immun. 69, 353359.
Meirelles, M.N., Juliano, L., Carmona, E., Silva, S.G., Costa, E.M., Murta,
A.C., Scharfstein, J., 1992. Inhibitors of the major cysteinyl proteinase
(gp57/51) impair host cell invasion and arrest the intracellular
development of Trypanosoma cruzi in vitro. Mol. Biochem. Parasitol.
52, 175184.
Moreno, S.N.J., Silva, J., Vercesi, A.E., Docampo, R., 1994. Cytosolic-free
calcium elevation in Trypanosoma cruzi is required for cell invasion.
J. Exp. Med. 180, 1535 1540.
Neira, I., Ferreira, A.T., Yoshida, N., 2002. Activation of distinct signal
transduction pathways in Trypanosoma cruzi isolates with differential
capacity to invade host cells. Int. J. Parasitol. 32, 405 414.
Neira, I., Silva, F.A., Cortez, M., Yoshida, N., 2003. Involvement of
Trypanosoma cruzi metacyclic trypomastigote surface molecule gp82
in adhesion to gastric mucin and invasion of epithelial cells. Infect.
Immun. 71, 557561.
Paiva, C.N., Souto-Padron, T., Costa, D.A., Gattass, C.R., 1998. High
expression of a functional cruzipain by a non-infective and non-
pathogenic Trypanosoma cruzi clone. Parasitology 117, 483490.
Ramirez, M.I., Ruiz, R.C., Araya, J.E., Franco da Silveira, J., Yoshida, N.,
1993. Involvement of the stage-specific 82-kilodalton adhesion
Table 1
Activation of distinct signal transduction pathways in metacyclic forms of
Trypanosoma cruzi CL isolate and clone CL-14 during host cell invasion
Parasite component
involved
Isolate CL Clone CL-14
Surface glycoprotein gp82 gp35/50
Protein tyrosine kinase Yes No
Phospholipase C Yes No
Intracellular Ca
2þ
store Inositol 1,4,5-triphosphate-
sensitive compartment
Acidocalcisome
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860 859
molecule of Trypanosoma cruzi metacyclic trypomastigotes in host cell
invasion. Infect. Immun. 61, 36363641.
Ruiz, R.C., Favoreto, S. Jr., Dorta, M.L., Oshiro, M.E.M., Ferreira,
A.T., Manque, P.M., Yoshida, N., 1998. Infectivity of Trypanosoma
cruzi isolates is associated with differential expression of surface
glycoproteins with differential Ca
2þ
signalling activity. Biochem. J.
330, 505 511.
Schenkman, S., Jiang, M., Hart, G.W., Nussenzweig, V., 1991. A novel
cell surface trans-sialidase of Trypanosoma cruzi generates a stage-
specific epitope required for invasion of mammalian cells. Cell 65,
11171125.
Schenkman, S., Pontes de Carvalho, L., Nussenzweig, V., 1992.
Trypanosoma cruzi trans-sialidase and neuraminidase
activities can be mediated by the same enzymes. J. Exp. Med.
175, 567 575.
Schenkman, S., Ferguson, M.A.J., Heise, N., Cardoso de Almeida, M.L.,
Mortara, R.A., Yoshida, N., 1993. Mucin-like glycoproteins linked to
the membrane by glycosylphosphatidylinositol anchor are the major
acceptors of sialic acid in a reaction catalysed by trans-sialidase in
metacyclic forms of Trypanosoma cruzi. Mol. Biochem. Parasitol. 59,
293304.
Tardieux, I., Nathanson, M.H., Andrews, N.W., 1994. Role in host cell
invasion of Trypanosoma cruzi-induced cytosolic free Ca
2þ
transients.
J. Exp. Med. 179, 1017 1022.
Teixeira, M.M.G., Yoshida, N., 1986. Stage-specific surface antigens
of metacyclic trypomastigotes of Trypanosoma cruzi identified by
monoclonal antibodies. Mol. Biochem. Parasitol. 18, 271282.
Yoshida, N., 1983. Surface antigens of metacyclic trypomastigotes of
Trypanosoma cruzi. Infect. Immun. 40, 836839.
Yoshida, N., Mortara, R.A., Araguth, M.F., Gonzalez, J., Russo, M., 1989.
Metacyclic neutralizing effect of monoclonal antibody 10D8 directed to
the 35- and 50-kilodalton surface glycoconjugates of Trypanosoma
cruzi. Infect. Immun. 57, 1663 1667.
Yoshida, N., Dorta, M.L., Ferreira, A.T., Oshiro, M.E.M., Mortara, R.A.,
Acosta-Serrano, A., Favoreto, S. Jr., 1997. Removal of sialic acid from
mucin-like surface molecules of Trypanosoma cruzi metacyclic
trypomastigotes enhances parasite host cell interaction. Mol. Biochem.
Parasitol. 84, 57 67.
Yoshida, N., Favoreto, S. Jr., Ferreira, A.T., Manque, P.M., 2000. Signal
transduction induced in Trypanosoma cruzi metacyclic trypomastigotes
during the invasion of mammalian cells. Braz. J. Med. Biol. Res. 33,
269278.
V.D. Atayde et al. / International Journal for Parasitology 34 (2004) 851–860860
INFECTION AND IMMUNITY, July 2007, p. 3264–3270 Vol. 75, No. 7
0019-9567/07/$08.00ϩ0 doi:10.1128/IAI.00262-07
Copyright © 2007, American Society for Microbiology. All Rights Reserved.
Expression and Cellular Localization of Molecules of the gp82 Family
in Trypanosoma cruzi Metacyclic Trypomastigotes
Vanessa D. Atayde, Mauro Cortez, Renata Souza, Jose´ Franco da Silveira, and Nobuko Yoshida*
Departamento de Microbiologia, Imunologia e Parasitologia, Universidade Federal de Sa˜o Paulo,
Rua Botucatu 862-6° andar, 04023-062, Sa˜o Paulo, Brazil
Received 16 February 2007/Returned for modification 22 March 2007/Accepted 6 April 2007
A member of the Trypanosoma cruzi gp82 family, expressed on metacyclic trypomastigote surface and
identified by monoclonal antibody (MAb) 3F6, plays a key role in host cell invasion. Apart from the gp82
defined by MAb 3F6, no information is available on members of this protein family. From cDNA clones
encoding gp82 proteins sharing 59.1% sequence identity, we produced the recombinant proteins J18 and C03,
the former containing and the latter lacking the epitope for MAb 3F6. Polyclonal antibodies to J18 and C03
proteins were generated and used, along with MAb 3F6, to analyze the expression and cellular localization of
gp82 family members in metacyclic forms of CL and G strains, which belong to highly divergent genetic groups.
By two-dimensional gel electrophoresis and immunoblotting, molecules of 82 to 86 kDa, focusing at pH 4.6 to
5.4, and molecules of 72 to 88 kDa, focusing at pH 4.9 to 5.7, were visualized in CL and G strains, respectively.
Flow cytometry and microscopic analysis revealed in both strains similar expression of MAb 3F6-reactive gp82
in live and permeabilized parasites, indicating its surface localization. The reaction of live parasites of both
strains with anti-J18 antibodies was weaker than with MAb 3F6 and was increased by permeabilization.
Anti-C03 antibodies bound predominantly to flagellar components in permeabilized G strain parasites, but in
the CL strain the flagellum was not the preferential target for these antibodies. Host cell invasion of metacyclic
forms was inhibited by J18 protein, as well as by MAb 3F6 and anti-J18 antibodies, but not by C03 protein or
anti-C03 antibodies.
Metacyclic trypomastigotes of different Trypanosoma cruzi
strains may engage different surface molecules to invade host
cells (32). In the highly infective CL strain, the metacyclic
stage-specific surface glycoprotein gp82 identified by monoclo-
nal antibody (MAb) 3F6 promotes target cell invasion by trig-
gering bidirectional signaling cascades leading to Ca
2ϩ
mobi-
lization in both parasite and target cells (17, 22, 34), which is an
event essential for parasite internalization (12, 26). Binding of
gp82 to target cells induces a Ca
2ϩ
-dependent disruption of
actin microfilaments (10), a process reported to facilitate par-
asite entry (21). gp82 also has the ability to bind to gastric
mucin (18), and this is crucial for the establishment of T. cruzi
infection by the oral route since the binding to mucin repre-
sents the first step toward the invasion of gastric mucosal
epithelium (8, 9). The poorly invasive G strain metacyclic
forms express MAb 3F6-reactive gp82 molecules, but they
preferentially use the mucin-like surface glycoproteins gp35/50
to enter host cells (22, 24, 33).
The MAb 3F6-reactive gp82 molecule is a member of a
multigene family, which is part of the large trans-sialidade/gp85
gene family (1). According to T. cruzi proteome analysis, 30 of
the 50 top-scoring proteins detected exclusively in the infective
trypomastigote forms are trans-sialidase (TS) family members
(2). The repertoire of metacyclic trypomastigote gp82 mole-
cules, the degree to which they are heterogeneous, the expres-
sion of members lacking the MAb 3F6 epitope, and their
presence in locations other than on the parasite surface are
matters that remain to be determined. The gp82 molecules
characterized so far are highly conserved in T. cruzi strains CL
and G, which belong to highly divergent genetic groups (5),
and show 97.9% peptide sequence identity overall and 100%
identity with regard to the cell binding site and the epitope for
MAb 3F6 (32).
In this study we isolated and characterized a new member of
the gp82 family and performed a global analysis on the expres-
sion as well as the cellular localization of gp82 proteins in
metacyclic forms of the CL and G strains. The strategy con-
sisted of the following steps: (i) isolation of a cDNA clone
encoding a member of the gp82 family lacking the epitope for
MAb 3F6, (ii) production of recombinant proteins with and
without the MAb 3F6 epitope, (iii) generation of antibodies
against the referred recombinant proteins, and (iv) two-dimen-
sional (2D) gel electrophoresis of metacyclic trypomastigote
extracts and immunoblotting in parallel with analysis by flow
cytometry and fluorescent microscope visualization of live as
well as permeabilized parasites, using MAb 3F6 and anti-gp82
polyclonal antibodies.
MATERIALS AND METHODS
Parasites. The following T. cruzi strains were used: CL, isolated from the
insect Triatoma infestans in the state of Rio Grande do Sul (4), and G, isolated
from an opossum in the Amazon (31). Parasites were maintained cyclically in
mice and in liver infusion tryptose. Before purification, in some cases the para-
sites were grown in Grace’s medium. Metacyclic forms from cultures in liver
infusion tryptose or Grace’s medium at the stationary growth phase were purified
by passage through a DEAE-cellulose column, as described previously (27).
Purification of RNA, RT-PCR, and cloning in pGEM-T. Purified CL strain
metacyclic trypomastigotes (1 ϫ 10
8
) were lysed with 1 ml of Trizol reagent
(Invitrogen). Following complete dissolution and the addition of 0.2 ml of chlo-
* Corresponding author. Mailing address: Escola Paulista de Medi-
cina, Universidade Federal de Sa˜o Paulo, Rua Botucatu 862-6° andar,
04023-062 Sa˜o Paulo, S.P., Brazil. Phone: 55 11 5576 4532. Fax: 55 11
5571 1095. E-mail: [email protected].
Published ahead of print on 16 April 2007.
3264
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
roform, the parasite preparation was centrifuged at 14,000 ϫ g for 15 min at 4°C.
The aqueous phase was collected, and an equal volume of isopropyl alcohol was
added to precipitate the total RNA. After washing with 75% ethanol, the pre-
cipitate was resuspended in RNase-free water. Reverse transcription-PCR (RT-
PCR) was performed using the Access Quick RT-PCR system (Promega). The
reaction mixture contained the following: total T. cruzi RNA; deoxynucleoside
triphosphate mix; avian myeloblastosis virus reverse transcriptase; Taq polymer-
ase; the forward primer SL (5Ј-GATACAGTTTCTGTACTATATTGAG-3Ј),
which is specific for the spliced leader sequence (GenBank accession no.
M30787); and the reverse primer 1775R (5Ј-GTTCCATTCGAAAGCATCCA
GTT-3Ј), which is specific for a sequence of the 3Ј region of the gp82 gene
(GenBank accession no. L14824). Production of cDNA was carried out at 48°C
for 45 min. Amplification was performed by running 45 cycles of denaturing,
annealing, and elongation at 94°C for 15 s, 44°C for 30 s, and 72°C for 1 min,
respectively. After purification with a GenClean kit (Bio 101), the PCR product
was cloned in the plasmid vector pGEM using a pGEM-T Easy Vector kit
(Promega). Following ligation to the vector, the product was transformed in
Escherichia coli strain DH5, and the colonies were grown in LB broth. Inserts
released from cDNA clones were screened with a
32
P-labeled probe containing
the full-length T. cruzi gp82 gene. The selected clones were sequenced using a Big
Dye terminator cycle sequencing ready reaction kit (Perkin-Elmer).
Production and purification of recombinant proteins J18 and C03. The re-
combinant protein J18, containing the full-length T. cruzi gp82 in frame with
glutathione S-transferase, was produced in E. coli DH5 by transforming the
bacteria with a pGEX-3 construct comprising the gp82 gene (23). All steps for
induction of the recombinant protein J18 and its purification are detailed else-
where (10). The recombinant protein C03, containing the full-length gp82 in
fusion with six histidine residues, was produced in E. coli BL21(DE3) by trans-
forming the bacteria with the construct pHIS-C03. This construct was generated
by PCR using the following primers: one containing the third ATG initiation
codon plus an artificial BamHI site (5Ј-ATTGGATCCGATGTGCTGCGCCA
CC-3Ј) and the other containing the stop codon plus an artificial HindIII site
(5Ј-GGAAGCTTTCTCAGTAAAGGGCCGC-3Ј). As a template, we used the
plasmid pGEM-T-C03. Following cloning in the vector pET-22(bϩ) (pHIS;
Novagen, Madison, WI), the clones were digested with BamHI/HindIII in order
to release the insert. Upon confirmation of the sequence, the clone pHIS-C03
was selected for further characterization. The transformed bacteria were grown
in LB medium and induced with 1 mM isopropyl--
D-thiogalactopyranoside for
4 h at 37°C, treated with 15 mg of lysozyme in phosphate-buffered saline (PBS)
for 30 min at room temperature, and then sonicated for 20 min and centrifuged
at 12,000 ϫ g for 30 min a 4°C. After three washings with 10 ml of CHAPS
(3-[(3-cholamidopropyl)-dimethylammonio]-1-propanesulfonate) buffer at 1%,
the precipitate was collected, resuspended in binding buffer (20 mM Na
2
HPO
4
,
0.5 M NaCl, 20 mM imidazole, 8 mM urea, pH 7.4), and incubated at room
temperature under agitation for 2 h. Thereafter, the sample in binding buffer was
centrifuged at 2,600 ϫ g for 10 min, and the supernatant was added to a 5-ml
Ni
3ϩ
column (Ni Sepharose 6 Fast Flow; Amersham Biosciences). Following
incubation for 30 min under agitation at room temperature, the bound protein
containing the histidine tail was eluted with the elution buffer (20 mM Na
2
HPO
4
,
0.5 M NaCl, 250 mM imidazole, 8 M urea, pH 7.4) and dialyzed against double-
distilled water for 48 h at 4°C. The amount of purified protein was quantified by
reaction with Coomassie Plus (Pierce) in 96-well plates, and readings were taken
at 620 nm. To certify that the desired protein was obtained, the purified samples
were analyzed in sodium dodecyl sulfate-polyacrylamide gel electrophoresis
(SDS-PAGE) gels stained with Coomassie blue and by immunoblotting using
anti-His antibodies (Amersham Biosciences).
Generation of antibodies to recombinant proteins J18 and C03. BALB/c mice
were injected by intraperitoneal route with recombinant protein (5 g/mouse) in
the presence of aluminum hydroxide (0.5 mg/mouse) as adjuvant. Fourteen days
later, all animals received the same amount of antigen plus adjuvant. Thereafter,
at 1-week intervals mice were given two more doses. Ten days after the last
immunizing dose, the mice were bled, and sera were collected and stored at
Ϫ20°C until used.
Isoelectric focusing, SDS-PAGE, and immunoblotting. The standard Western
blot analysis was performed, as previously described (33), by applying NP-40-
solubilized parasite extracts corresponding to 3 ϫ 10
7
cells into each well of 10%
SDS-PAGE gels. For 2D electrophoresis, samples containing 5 ϫ 10
8
metacyclic
trypomastigotes were washed three times in 25 mM HEPES, pH 7.4, containing
0.9% NaCl; samples were boiled for 5 min in 0.2% SDS lysis buffer and main-
tained on ice before incubation with 2D buffer (7 M urea, 2 M thiourea, 1%
dithiothreitol [DTT], 2% Triton X-100, 0.5% immobilized pH gradient buffer,
pH 4 to 7) containing protease inhibitors (100 M phenylmethylsulfonyl fluoride,
1 M pepstatin, 100 M leupeptin, and 5 mM EDTA) for 30 min at room
temperature. After centrifugation at 23,100 ϫ g for 10 min, the supernatant was
applied to immobilized pH gradient gel strips with 0.5% (pH 4.0 to 7.0) ampho-
lites (Amersham Biosciences). Isoelectric focusing was performed in an Ettan
IPGphor system (Amersham Biosciences) by applying 500 V, 1,000 V, 8,000 V,
and 7,400 V, sequentially at 1-h intervals. Thereafter, the gel strips were incu-
bated for 15 min in a solution containing 6 M urea, 50 mM Tris, pH 8.8, 30%
glycerol, 2% SDS, and 25 mM DTT. Following another 15-min incubation in this
solution without DTT but containing 125 mM iodoacetamide plus 0.02% bro-
mophenol blue, the samples were subjected to electrophoresis in 10% SDS-
PAGE gels. The proteins were then transferred to a nitrocellulose membrane,
which was processed for reaction with anti-gp82 antibodies.
Flow cytometry. Live metacyclic trypomastigotes (3 ϫ 10
7
) were incubated for
1 h on ice with anti-gp82 antibodies. After washings in PBS and fixation with 2%
paraformaldehyde for 30 min, the parasites were incubated with anti-mouse
immunoglobulin G (IgG) conjugated to fluorescein at room temperature for 1 h.
Following two more washes, the number of fluorescent parasites was estimated
with a Becton Dickinson FACscan cytometer. Assays with fixed and permeabil-
ized parasites were carried out as follows: fixation with 2% paraformaldehyde for
30 min, washings in PBS, treatment with 0.1% saponin in PBS at room temper-
ature for 30 min, washings in PBS, and incubation with antibodies as described
above.
Microscopic visualization of fluorescent parasites. Live metacyclic forms were
incubated for1honicewith anti-gp82 antibodies, washed, fixed with 3.5%
formaldehyde in PBS, and placed onto glass slides and dried. Afterwards, the
parasites were incubated sequentially with fluorescein-conjugated anti-mouse
IgG diluted 1:40 in PGN (0.15% gelatin in PBS containing 0.1% sodium azide)
for 1 h, and 10 M DAPI (4Ј,6Ј-diamidino-2-phenylindole; Molecular Probes)
for visualization of kinetoplast and nucleus. Images were acquired on a Nikon
E600 fluorescence microscope coupled to a Nikon DXM 1200F digital camera
using ACT-1 software. In parallel, the parasites were first fixed with 3.5% form-
aldehyde, washed, and then processed as above, except that a 1-h incubation with
anti-gp82 antibodies was carried out in the presence of 0.1% saponin in PGN for
parasite permeabilization.
Host cell invasion assay. HeLa cells, the human carcinoma-derived epithelial
cells, were grown at 37°C in Dulbecco’s minimum essential medium, supple-
mented with 10% fetal calf serum, streptomycin (100 g/ml), and penicillin (100
U/ml) in a humidified 5% CO
2
atmosphere. Cell invasion assays were carried out
as detailed elsewhere (33) by seeding the parasites onto each well of 24-well
plates containing 13-mm-diameter round glass coverslips coated with 1.5 ϫ 10
5
HeLa cells. After a 1-h incubation with parasites at a parasite:cell ratio of 10:1,
the coverslips were washed in PBS and stained with Giemsa, and the numbers of
intracellular parasites were counted.
Nucleotide sequence accession number. The sequence of the recombinant
protein C03 has been deposited in the GenBank database under accession
number EF445668.
RESULTS
Isolation of a new member of the gp82 family. We cloned a
full-length gp82 cDNA (C03) by RT-PCR, using a set of prim-
ers based on the sequence of clone J18, which codes for a gp82
protein containing the epitope for MAb 3F6 (1) and the
miniexon sequence (29) present in all trypanosomatid mRNAs.
Translation of C03 cDNA in each of the six possible reading
frames indicated only one large open reading frame with three
in-frame ATG initiator codons. Within this open reading
frame the first potential start codon is separated from the
spliced leader sequence by 89 bp. The second and third start
codons are in the same reading frame as the first ATG and they
are 51 and 159 bp downstream from it, respectively. Analysis of
the 5Ј sequences flanking the three potential start codons in-
dicates that the third ATG best fits the Kozak eukaryotic
consensus sequences (15). This argues for the use of the third
ATG as the initiating methionine. However, we cannot exclude
the possibility of translation from the first or the second in-
frame methionine, provided that the amino acid sequence of
the amino-terminal end of the native gp82 has not yet been
determined. Assuming the third codon as the initiating methi-
VOL. 75, 2007 EXPRESSION OF T. CRUZI gp82 FAMILY PROTEINS 3265
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
onine, C03 cDNA encodes a polypeptide of 668 amino acids
with a predicted molecular mass of 72.54 kDa and pI 5.03 (Fig.
1). It has two highly conserved copies of the sialidase motif
SXDXGXTW (Asp box) and a complete copy of the subter-
minal motif VTVKNVFLYNR characteristic of all members of
the TS superfamily (13). Although C03 does not display an
N-terminal sequence predicted to be a signal peptide by the
SignalP 3.0 program (3), it has a typical glycosylphosphatidyl-
inositol anchor signal sequence at the carboxy terminus (Fig. 1).
The amino acid sequence deduced from cDNA clone C03
FIG. 1. Sequences of proteins of T. cruzi metacyclic trypomastigote gp82 family members. Shown are the proteins C03 (CL strain), R31 (CL
strain), and J18 (G strain). In proteins R31 and J18 are the sequences (indicated by lines above the sequences) identified as the epitope for MAb
3F6 (P3) and as the conformational cell binding site (P4 and P8). The alternative initiator methionines in the C03 are indicated by asterisks. The
conserved motif VTV, characteristic of all members of the gp85/TS family, and the potential glycosylphosphatidylinositol (GPI) anchor consensus
sequence are indicated (lines above sequences). Points represent residues that are conserved in the three proteins, nonconserved amino acids are
indicated, and dashes represent residues that are lacking in the indicated sequence. Note the high degree of identity between protein R31 and J18,
as opposed to the much lower sequence identity of these proteins in relation to protein C03.
3266 ATAYDE ET AL. I
NFECT.IMMUN.
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
displayed considerable differences from the sequences de-
duced from clones J18 and R31 (GenBank accession no.
AF128843) isolated in previous studies from cDNA libraries
from G and CL strain metacyclic trypomastigotes, respectively.
The overall amino acid sequence identity was 59.1% between
C03 and J18 proteins and 49% between C03 and R31 proteins.
Comparative analysis of recombinant proteins C03 and J18.
Using synthetic peptides based on the J18 protein sequence,
the epitope for MAb 3F6 has been mapped, and it corresponds
to the sequence represented by peptide P3 (16). As shown in
Fig. 2A, compared to the sequence of J18 corresponding to P3,
the most conspicuous difference in the C03 protein was the
substitution of three uncharged residues by charged ones (Fig.
2A). Reflecting this difference, the C03 protein failed to be
recognized by MAb 3F6 (Fig. 2B). Considerable differences
were also seen in the sequence corresponding to the cell bind-
ing site, which in the J18 protein is formed by juxtaposition of
two sequences, P4 and P8, containing several charged residues
separated by a more hydrophobic stretch (16); comparison of
the sequences of the two proteins revealed substitutions of
charged residues for uncharged amino acids and vice versa
(Fig. 2A). Accordingly, the binding capacity of the C03 protein
to HeLa cells was lower than that of the J18 protein (Fig. 2C).
Expression and cellular localization of molecules of the
gp82 family in G and CL strain metacyclic forms. We gener-
ated antibodies against C03 and J18 recombinant proteins by
immunizing groups of mice with either of these proteins. The
anti-C03 antibody also reacted with J18 protein, and anti-J18
antibodies recognized C03 protein as well, and in both cases
the reaction was stronger with the protein that served as im-
munogen (Fig. 3). Both antibodies reacted similarly with gp82
molecules from metacyclic trypomastigotes of G and CL
strains (Fig. 3). Immunoblots of CL strain parasite extracts
subjected to 2D gel electrophoresis showed the MAb 3F6-
reactive gp82 molecules focusing at a pH range of 4.6 to 5.1
(data not shown). We confirmed the presence of these gp82
molecules on the parasite surface by flow cytometry analysis
and observation with fluorescence microscopy, which revealed
no difference in the reactivity of live and permeabilized para-
sites with MAb 3F6 (Fig. 4). As regards the gp82 molecules
recognized by anti-J18 or anti-C03 antibodies, with isoelectric
points ranging from pH 4.6 to 5.4 (data not shown), their
location was at least in part intracellular, as deduced from the
high reactivity of permeabilized parasites compared to their
FIG. 2. Properties of recombinant proteins J18 and C03. (A) The
sequences identified as the epitope for MAb 3F6 (P3) and as the con-
formational cell binding site (P4 and P8) are aligned to show the
differences between J18 and C03 proteins, with asterisks indicating the
changed residues. (B) Immunoblot of purified J18 and C03 proteins
using MAb 3F6. Note the lack of reactivity of C03 protein. (C) Binding
of proteins J18 and C03 to HeLa cells, assayed as described in Mate-
rials and Methods. The reaction was revealed by sequential incubation
with polyclonal antiserum directed to J18 or to C03 and anti-mouse
IgG conjugated to peroxidase. OD
492
, optical density at 492 nm.
FIG. 3. Cross-reactivity between anti-J18 and anti-C03 antibodies.
Purified recombinant proteins J18 and C03, as well as the metacyclic
trypomastigote extract of T. cruzi strains G and CL, were analyzed by
immunoblotting using polyclonal antibodies to J18 or C03 protein.
FIG. 4. Cellular localization of proteins of the gp82 family in meta-
cyclic forms of the T. cruzi CL strain. Live or permeabilized parasites
were reacted with MAb 3F6 or polyclonal antibodies to J18 or C03 and
processed for analysis by flow cytometry or for visualization by fluo-
rescence microscopy. DAPI staining of parasite kinetoplast and nu-
cleus appears in blue. Note the similar expression patterns of MAb
3F6-reactive molecules in live and permeabilized parasites, whereas
the molecules recognized by anti-J18 and anti-C03 antibodies were in
part localized intracellularly, as indicated by the high fluorescence inten-
sity of permeabilized parasites compared to their live counterparts.
V
OL. 75, 2007 EXPRESSION OF T. CRUZI gp82 FAMILY PROTEINS 3267
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
live counterparts, and this was clearly visible by observations
with fluorescence microscopy (Fig. 4). We also noted that in
live parasites the distribution of molecules recognized by anti-
C03 antibodies was not homogeneous, particularly in the pos-
terior end, where they were more concentrated at some spots
(Fig. 4).
In G strain metacyclic forms, 2D gel electrophoresis and
immunoblotting revealed a similar profile of gp82 bands focus-
ing at pH 4.9 to 5.7 when MAb 3F6 or anti-J18 antibodies were
used (data not shown). As in the CL strain, the MAb 3F6-
reactive gp82 molecules were on the surface, and no difference
in reactivity was observed between live and permeabilized par-
asites (Fig. 5). On the other hand, a portion of gp82 molecules
reacting with anti-J18 antibodies was intracellularly located, as
indicated by the considerably increased reactivity when the
parasites were permeabilized (Fig. 5). The most striking ob-
servation in the G strain was the reaction of anti-C03 antibod-
ies with permeabilized metacyclic forms. These antibodies,
which reacted with gp82 bands focusing at pH 4.9 to 5.5 (data
not shown), barely recognized live parasites but reacted in-
tensely with flagellar components of permeabilized cells (Fig.
5), which means that these are intraflagellar molecules. In
order to determine whether the flagellar location of proteins
recognized by anti-C03 antibodies was a characteristic of
strains of group 1 T. cruzi, to which the G strain belongs (5),
two other strains of this group were examined. The same pro-
file of the G strain, with the predominant reaction with flagella,
was revealed in permeabilized metacyclic forms of these strains
(data not shown).
Effect of different proteins and antibodies on host cell inva-
sion by metacyclic trypomastigotes. To test the effect of pro-
teins and antibodies generated in this study, we used CL strain
metacyclic forms and epithelial HeLa cells. Parasites were in-
cubated with HeLa cells in the absence or presence of purified
recombinant protein, J18, C03, or glutathione S-transferase, at
40 g/ml for 1 h. In contrast to the J18 protein, which inhibited
parasite internalization by ϳ65%, C03 protein had no inhibi-
tory effect (Fig. 6A). Why the C03 protein does not inhibit T.
cruzi invasion even though it binds to host cells, albeit to a
lesser extent than J18 (Fig. 2C), is not known. One possibility
is that the C03 protein binds to a target receptor distinct from
the J18 receptor, which recognizes a sequence different from
that corresponding to P4/P8. We also examined the effect of
anti-C03 and anti-J18 antibodies on HeLa cell invasion. Para-
sites were incubated for 30 min with anti-J18 or anti-C03 an-
tiserum diluted 1:10 or with ascitic fluid containing MAb 3F6
diluted 1:10. After the antibody was washed out, the parasites
were seeded onto HeLa cells. As shown in Fig. 6B, parasite
entry was inhibited ϳ50% by MAb 3F6 and ϳ35% by anti-J18
antibodies, whereas anti-C03 antibodies showed no effect.
FIG. 5. Cellular localization of proteins of the gp82 family in meta-
cyclic forms of the T. cruzi G strain. Live or permeabilized parasites
were reacted with MAb 3F6 or polyclonal antibodies to J18 or C03 and
processed for analysis by flow cytometry or for visualization by fluo-
rescence microscopy. Note the similar expression patterns of MAb
3F6-reactive molecules in live and permeabilized parasites, whereas
the molecules recognized by anti-J18 antibodies were in part located
intracellularly, and those reacting with anti-C03 antibodies were asso-
ciated with the flagellum and did not react with live parasites.
FIG. 6. Effect of different proteins and antibodies on host cell in-
vasion by T. cruzi metacyclic trypomastigotes (CL strain). (A) HeLa
cells were incubated with parasites in the absence or presence of the
indicated protein at 40 g/ml. (B) Parasites were incubated with the in-
dicated antibodies for 30 min. After antibodies were washed out, the
parasites were incubated with HeLa cells. In both panels, the incuba-
tion time was 1 h, the cells were stained with Giemsa, and the numbers
of intracellular parasites were counted in at least 500 cells. Values in
panel A are the means Ϯ standard deviation of three independent
assays performed in duplicate. The difference between the control and
the cells incubated with J18 protein was significant (P Ͻ 0.0005). In
panel B the values are the means Ϯ standard deviation of triplicates of
one representative experiment. A significant difference between the
control and the parasites incubated with MAb 3F6 (P Ͻ 0.0005) or
with anti-J18 antibodies (P Ͻ 0.005) was found.
3268 ATAYDE ET AL. I
NFECT.IMMUN.
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
DISCUSSION
Our study provides novel information on the expression and
cellular localization of gp82 family members in T. cruzi meta-
cyclic trypomastigotes. This was achieved through the isolation
of a cDNA clone encoding a member of the gp82 family dis-
tinct from the previously characterized cDNA clone selected
for its reactivity with MAb 3F6 (1). From these cDNA clones
encoding proteins with 59.1% identity, we generated recombi-
nant proteins C03 and J18 as well as the corresponding anti-
bodies (Fig. 1 and 3), which served to identify and localize the
gp82 molecules in the parasites.
As a common characteristic, the members of the gp82 family
recognized by MAb 3F6, anti-J18, or anti-C03 antibodies had
isoelectric points at the acidic pH range, with the CL and G
strain molecules focusing at pH 4.6 to 5.4 and at 4.9 to 5.7,
respectively. Otherwise, the gp82 molecules reacting with any
of these antibodies showed considerable differences, particu-
larly in their cellular localization. We confirmed that the MAb
3F6-reactive molecules are present predominantly on the sur-
face of metacyclic trypomastigotes of both the CL and G
strains (Fig. 4 and 5). In T. cruzi strains that enter host cells in
a gp82-mediated manner, the MAb 3F6-reactive molecules
contribute to the invasion process (20), primarily by binding to
the host cell surface molecules and triggering the activation of
signal transduction pathways in both the parasite and target
cells (32). On the other hand, the gp82 molecules recognized
by anti-J18 antibodies appear to be expressed at low levels on
the parasite surface and therefore may have a more marginal
role, if any, in parasite internalization. In CL strain metacyclic
trypomastigotes, which engage surface gp82 molecules to in-
vade target cells (20, 22), the intensity of the reaction of live
parasites with anti-J18 antibodies was much reduced compared
to that with MAb 3F6 (Fig. 4). If MAb 3F6-reactive compo-
nents are included in the gp82 repertoire identified by anti-J18
antibodies, then the gp82 population present on the surface
that is specifically recognized by these polyclonal antibodies
may be even smaller. Although to a lesser degree, the intensity
of reaction of live metacyclic forms of G strain with anti-J18
antibodies was also found to be weaker than with MAb 3F6,
but in permeabilized parasites the intensity of reaction was
comparable to that with the MAb (Fig. 5). Taking these results
together, we provide experimental evidence for the existence
of several subpopulations among the gp82 metacyclic trypo-
mastigote proteins, including those that are not targeted to the
cell membrane.
What was particularly striking and unpredicted was the find-
ing that members of the gp82 family may be localized to the
flagellum in T. cruzi in a strain-dependent manner. That the
profile of reactivity of metacyclic trypomastigotes with anti-
C03 antibodies could be quite different from that with anti-J18
antibodies was expected, as they were elicited by proteins with
considerable sequence diversity (Fig. 1). However, that gp82
family members could have such different cellular localizations
and that this distribution varied depending on the T. cruzi
strain (Fig. 4 and 5) came as a surprise. Anti-C03 antibodies
did not react with live G strain metacyclic forms but reacted
strongly with components associated with the flagellum in per-
meabilized parasites (Fig. 5), indicating that they recognize
intraflagellar molecules. In the CL strain, the profile of the
reaction of live as well as permeabilized parasites with anti-C03
antibodies was similar to that with anti-J18 antibodies, with no
preferential association with flagellum being observed (Fig. 4).
Several T. cruzi molecules associated to flagellum have been
previously described, among them a group of surface proteins
of the 160-kDa family, belonging to the TS/gp85 superfamily
and sharing 48% identity with gp85 proteins (25, 30). These
proteins were detected on the cell membrane in the flagellar
pocket of bloodstream trypomastigotes (28, 30). Comparison
of C03 and Fl-160 amino acid sequences shows 29% identity,
suggesting that they could share epitopes. Other proteins as-
sociated with T. cruzi flagellum are the flagellar calcium-bind-
ing protein (6), the glycoprotein gp72 (19), and a high-molec-
ular-mass (300 kDa) protein built up mostly by nearly identical
repeats of 68 amino acids arranged in tandem (11). The flagel-
lar calcium-binding protein, which is a flagellum-specific cal-
cium sensor, is associated with the flagellar membrane via its
N-terminal myristate and palmitate moieties in a calcium-mod-
ulated, conformation-dependent manner (6). In epimastigotes,
gp72 is predominantly membrane associated and located on
the cell surface, being evenly distributed over the cell body and
somewhat concentrated in the proximal region of the flagellum
(14). The morphology of gp72-null mutants was found to be
dramatically different from wild-type parasites, and the normal
attachment of the flagellum to the cell membrane was lost (7).
What role is played by intraflagellar gp82 molecules recog-
nized by anti-C03 antibodies remains to be determined.
We also confirmed here the involvement of MAb 3F6-reac-
tive gp82 molecules in target cell invasion. Purified J18 protein,
as well as MAb 3F6 and anti-J18 antibodies, inhibited meta-
cyclic trypomastigote entry into host cells, whereas C03 protein
and anti-C03 antibodies showed no inhibitory effect, as ex-
pected (Fig. 6). This and the new data provided in this study
contribute to our knowledge of T. cruzi gp82 family members,
which may have functions other than interacting with target
cells. Additionally, the differential cellular localization of gp82
molecules recognized by anti-C03 antibodies in G and CL
strains (Fig. 4 and 5) further reminds us of the remarkable
differences between these two strains that have been described
in previous studies.
ACKNOWLEDGMENTS
This work was supported by Fundac¸a˜o de Amparo a` Pesquisa do
Estado de Sa˜o Paulo and Conselho Nacional de Desenvolvimento
Cientı´fico e Tecnolo´gico, Brazil.
REFERENCES
1. Araya, J. E., M. I. Cano, N. Yoshida, N., and J. Franco da Silveira. 1994.
Cloning and characterization of a gene for the stage-specific 82-kDa surface
antigen of metacyclic trypomastigotes of Trypanosoma cruzi. Mol. Biochem.
Parasitol. 65:161–169.
2. Atwood, J. A., D. B. Weatherly, T. A Minning, B. Bundy, C. Cavola, F. R.
Operdoes, R. Orlando, and R. L. Tarleton. 2005. The Trypanosoma cruzi
proteome. Science 309:473–476.
3. Bendtsen, J. D., H. Nielsen, G. von Heijne, and S. Brunak. 2004. Improved
prediction of signal peptides: SignalP 3.0. J. Mol. Biol. 340:783–795.
4. Brener, Z., and E. Chiari. 1963. Variac¸o˜es morfolo´gicas observadas em
diferentes amostras de Trypanosoma cruzi. Rev. Inst. Me´d. Trop. Sa˜o Paulo
5:220–224.
5. Briones, M. R. S., R. P. Souto, B. S. Stolf, and B. Zingales. 1999. The
evolution of two Trypanosoma cruzi subgroups inferred from rRNA genes
can be correlated with the interchange of American mammalian faunas in
the Cenozoic and has implications to pathogenicity and host specificity. Mol.
Biochem. Parasitol. 104:219–232.
6. Buchanan, K. T., J. B. Ames, S. H. Asfaw, J. N. Wingard, C. L. Olson, P. T.
VOL. 75, 2007 EXPRESSION OF T. CRUZI gp82 FAMILY PROTEINS 3269
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
Campana, A. P. Arau´jo, and D. M. Engman. 2005. A flagellum-specific
calcium sensor. J. Biol. Chem. 280:40104–40111.
7. Cooper R., A. R. de Jesus, and G. A. Cross. 1993. Deletion of an immuno-
dominant Trypanosoma cruzi surface glycoprotein disrupts flagellum-cell ad-
hesion. J. Cell Biol. 122:149–156.
8. Cortez, M., I. Neira, D. Ferreira, A. O. Luquetti, A. Rassi, and N. Yoshida.
2003. Infection by Trypanosoma cruzi metacyclic forms deficient in gp82 but
expressing a related surface molecule, gp30. Infect. Immun. 71:6184–6191.
9. Cortez, M., M. R. Silva, I. Neira, D. Ferreira, G. R. S. Sasso, A. O. Luquetti,
A. Rassi, and N. Yoshida. 2006. Trypanosoma cruzi surface molecule gp90
downregulates invasion of gastric mucosal epithelium in orally infected mice.
Microbes Infect. 8:36–44.
10. Cortez, M., V. Atayde, and N. Yoshida. 2006. Host cell invasion mediated by
Trypanosoma cruzi surface molecule gp82 is associated with F-actin disas-
sembly and is inhibited by enteroinvasive Escherichia coli. Microbes Infect.
8:1502–1512.
11. Cotrim, P. C., G. Paranhos-Baccala, M. R. Santos, C. Mertensen, M. I.
Cano, M. Jolivet, M. E. Camargo, R. A. Mortara, and J. F. Silveira. 1995.
Organization and expression of the gene encoding an immunodominant
repetitive antigen associated to the cytoskeleton of Trypanosoma cruzi. Mol.
Biochem. Parasitol. 71:89–98.
12. Docampo, R., and S. J. Moreno. 1996. The role of Ca2
ϩ
in the process of cell
invasion by intracellular parasites. Parasitol. Today 12:61–65.
13. Frasch, A. A. C. 2000. Functional diversity in the trans-sialidase and mucin
families in Trypanosoma cruzi. Parasitol. Today 16:282–286.
14. Haynes, P. A., D. G. Russell, and G. A. Cross. 1996. Subcellular localization
of Trypanosoma cruzi glycoprotein gp72. J. Cell Sci. 109:2979–2988.
15. Kozak, M. 1989. The scanning model for translation: an update. J. Cell Biol.
108:229–241.
16. Manque, P. M., D. Eichinger, M. A. Juliano, L. Juliano, J. Araya, and N.
Yoshida. 2000. Characterization of the cell adhesion site of Trypanosoma
cruzi metacyclic stage surface glycoprotein gp82. Infect. Immun. 68:478–484.
17. Neira, I., A. T. Ferreira, and N. Yoshida. 2002. Activation of distinct signal
transduction pathways in Trypanosoma cruzi isolates with differential capac-
ity to invade host cells. Int. J. Parasitol. 32:405–414.
18. Neira, I., F. A. Silva, M. Cortez, and N. Yoshida. 2003. Involvement of
Trypanosoma cruzi metacyclic trypomastigote surface molecule gp82 in ad-
hesion to gastric mucin and invasion of epithelial cells. Infect. Immun.
71:557–561.
19. Nozaki, T., and G. A. Cross. 1994. Functional complementation of glyco-
protein 72 in a Trypanosoma cruzi glycoprotein 72 null mutant. Mol. Bio-
chem. Parasitol. 67:91–102.
20. Ramirez, M. I., R. C. Ruiz, J. E. Araya, J. Franco da Silveira, and N.
Yoshida. 1993. Involvement of the stage-specific 82-kilodalton adhesion mol-
ecule of Trypanosoma cruzi metacyclic trypomastigotes in host cell invasion.
Infect. Immun. 61:3636–3641.
21. Rodriguez, A., M. G. Rioult, A. Ora, and N. N. Andrews. 1995. A trypano-
some-soluble factor induces IP
3
formation, intracellular Ca
2ϩ
mobilization
and microfilament rearrangement in host cells. J. Cell Biol. 129:1263–1273.
22. Ruiz, R. C., S. Favoreto, M. L. Dorta, M. E. M. Oshiro, A. T. Ferreira, P. M.
Manque, and N. Yoshida. 1998. Infectivity of Trypanosoma cruzi strains is
associated with differential expression of surface glycoproteins with differ-
ential Ca
2ϩ
signaling activity. Biochem. J. 330:505–511.
23. Santori, F. R., M. L. Dorta, L. Juliano, M. A. Juliano, J. Franco da Silveira,
R. C. Ruiz, and N. Yoshida. 1996. Identification of a domain of Trypanosoma
cruzi metacyclic trypomastigote surface molecule gp82 required for attach-
ment and invasion of mammalian cells. Mol. Biochem. Parasitol. 78:209–216.
24. Schenkman, S., M. A. J. Ferguson, N. Heise, M. L. Cardoso de Almeida,
R. A. Mortara, and N. Yoshida. 1993. Mucin-like glycoproteins linked to the
membrane by glycosylphosphatidylinositol anchor are the major acceptors of
sialic acid in a reaction catalyzed by trans-sialidase in metacyclic forms of
Trypanosoma cruzi. Mol. Biochem. Parasitol. 59:293–304.
25. Takle, G. B., J. O’Connor, A. J. Young, and G. A. Cross. 1992. Sequence
homology and absence of mRNA defines a possible pseudogene member of
the gp85/sialidase multigene family. Mol. Biochem. Parasitol. 56:117–128.
26. Tardieux, I., M. H. Nathanson, and N. W. Andrews. 1994. Role in host cell
invasion of Trypanosoma cruzi-induced cytosolic free Ca
2ϩ
transients. J. Exp.
Med. 179:1017–1022.
27. Teixeira, M. M. G., and N. Yoshida. 1986. Stage-specific surface antigens of
metacyclic trypomastigotes of Trypanosoma cruzi identified by monoclonal
antibodies. Mol. Biochem. Parasitol. 18:271–282.
28. Van Voorhis, W. C., and H. Eisen. 1989. Fl-160. A surface antigen of
Trypanosoma cruzi that mimics mammalian nervous tissue. J. Exp. Med.
169:641–652.
29. Walder, J. A., P. S. Eder, D. M. Engman, S. T. Brentano, R. Y. Walder, D. S.
Knutzon, D. M. Dorfman, and J. E. Donelson. 1986. The 35-nucleotide
spliced leader sequence is common to all trypanosome messenger RNAs.
Science 233:569–571.
30. W. C. Van Voorhis, L. Barrett, R. Koelling, and A. G. Farr. 1993. FL-160
proteins of Trypanosoma cruzi are expressed from a multigene family and
contain two distinct epitopes that mimic nervous tissues. J. Exp. Med. 178:
681–694.
31. Yoshida, N. 1983. Surface antigens of metacyclic trypomastigotes of Trypano-
soma cruzi. Infect. Immun. 40:836–839.
32. Yoshida, N. 2006. Molecular basis of mammalian cell invasion of Trypano-
soma cruzi. An. Acad. Bras. Cieˆnc. 78:87–111.
33. Yoshida, N., R. A. Mortara, M. F. Araguth, J. C. Gonzalez, and M. Russo.
1989. Metacyclic neutralizing effect of monoclonal antibody 10D8 directed to
the 35- and 50-kilodalton surface glycoconjugates of Trypanosoma cruzi.
Infect. Immun. 57:1663–1667.
34. Yoshida, N., S. Favoreto, Jr., A. T. Ferreira, and P. M. Manque. 2000. Signal
transduction induced in Trypanosoma cruzi metacyclic trypomastigotes dur-
ing the invasion of mammalian cells. Braz. J. Med. Biol. Res. 33:269–278.
Editor: W. A. Petri, Jr.
3270 ATAYDE ET AL. INFECT.IMMUN.
at Sistema Integrado de Bibliotecas-USP/FOB on November 27, 2007 iai.asm.orgDownloaded from
172 Original article
A recombinant protein based on Trypanosoma cruzi
surface molecule gp82 induces apoptotic cell death in
melanoma cells
Vanessa D. Atayde
a
, Miriam G. Jasiulioni s
b
, Mauro Cortez
a
and
Nobuko Yoshida
a
Trypanosoma cruzi infection is known to confer resistance
to tumor development in mice, and in-vitro studies have
shown the toxic effects of parasite extracts on cancer
cell cultures. Investigations in which T. cruzi molecules
exhibit antitumor activity have just begun. Here, we used
a tumorigenic cell line Tm5, derived from mouse
melanocytes melan-a, to test the effect of J18, a
recombinant protein based on T. cruzi surface molecule
gp82 fused to glutathione-S-transferase (GST). J18
induced actin cytoskeleton disruption in Tm5 but not in
melan-a cells. Several changes indic ative of apoptosis
were detected in Tm5 melanoma cells but not in
melan-a cells treated with J18, such as the flipping of
phosphatidylserine from the inner to the external side
of the plasma membrane, altered nuclear morphology,
DNA fragmentation, increase in mitochondria
depolarization, and in caspase-3 activity. Retention of
NF-jB in the cytoplasm was another alteration observed
specifically in J18-treated Tm5 cells. No such alterations
were found in Tm5 cells treated with GST. In-vivo
experiments showed that C57BL/6 mice inoculated with
Tm5 cells, treated at the site of tumor cell inoculation with
J18, developed tumors of smaller size than mice treated
with phosphate-buffered saline or GST and survived
longer. Melanoma Res 18:172–183
c
2008 Wolters Kluwer
Health | Lippincott Williams & Wilkins.
Melanoma Research 2008, 18:172–183
Keywords: surfac e molecule gp82, melanoma cells, Trypanosoma cruzi
a
Department of Microbiology, Immunology and Parasitology and
b
Department of
Pharmacology, Federal University of Sa
˜
o Paulo, Sa
˜
o Paulo, Brasil
Correspondence to Nobuko Yoshida, Escola Paulista de Medicina,
Universidade Federal de Sa
˜
o Paulo, Rua Botucatu, 8 62-61 Andar, 04023-062,
Sa
˜
o Paulo, SP, Brasil
Tel: + 55 115 576 4532; fax: + 55 115 571 1095;
Received 9 October 2007 Accepted 27 February 2008
Introduction
The incidence of cutaneous malignant melanoma, the
most aggressive form of skin cancer, has increased in
recent decades in most fair-skinned people throughout
the world [1,2]. Melanoma arises from the malignant
transformation of melanocytes, the pigment-producing
cells that reside in the basal layer of the epidermis,
through a complex process involving genetic as well as
environmental factors. Failure to undergo apoptosis is
among the essential alterations that dictate malignant
growth, which include insensitivity to growth-inhibitory
signals, limitless replicative potential, sustained angio-
genesis, tissue invasion, and metastasis [3]. Strategies
to counterbalance this failure may therefore contribute to
anticancer therapies.
Apoptosis, a physiological process of cell death, is
triggered by signals from either the intracellular or
extracellular milieu and is orchestrated by a suicide
machinery conserved through evolution, in which a
proteolytic system involving a family of proteases known
as caspases is a core component [4]. DNA fragmentation,
exposure of phosphatidylserine in the outer leaflet of the
plasma membrane, and loss of mitochondrial membrane
permeability are among the alterations that accompany
apoptosis. From a number of studies in recent years, the
actin cytoskeleton has emerged as a key regulator of
apoptosis [5]. By preventing actin filament elongation
with cytochalasin D, apoptosis was induced in adherent
epithelial cells [6]. Treatment of cells with cytochalasin
D enhanced apoptosis in human leukemia CMK-7
cells by greatly accelerating caspase-3 activation [7].
Jasplakinolide, a drug that stabilizes and aggregates actin
filaments, was shown to increase apoptosis induced by
cytokine deprivation in CTLL-20 cells, the murine
interleukin 2-dependent T cell line, apparently by acting
upstream of the induction of capsase-3-like activity [8].
As regards the cell death induced by jasplakinolide,
through a caspase-3-like protease pathway, an important
finding was that Jurkat cells, derived from human acute
lymphoblastic T cell leukemia, were more susceptible
than normal nontransformed cells [9].
Resistance to tumors in infections by microbial patho-
gens, including the protozoan parasites Toxoplasma gondii
and Trypanosoma cruzi, has been reported [10,11]. In-vitro
experiments have shown the direct inhibitory effect of
T. cruzi lysates on cultured human breast cancer cells [12]
and sarcoma-180 cells [13]. These antitumor effects
may be because of the proapoptotic activity of T. cruzi
0960-8931
c
2008 Wolters Kluwer Health | Lippincott Williams & Wilkins
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
components. A T. cruzi-released protein Tc52, when in
fusion with glutathione-S-transferase (GST), exhibited
apoptosis-inducing effect on a human T-cell leukemic cell
line [14]. Other T. cruzi components with proapoptotic
activity include a ceramide-containing glycolipid [15]
and trans-sialidase [16,17]. The mechanisms by which
these parasite molecules induce apoptosis have not been
clarified.
Using a line of nontumorigenic mouse melanocytes,
melan-a [18], and the tumorigenic cell line Tm5, derived
from melan-a [19], we aimed in this study to investigate
the apoptosis-inducing potential of a recombinant pro-
tein, based on T. cruzi surface molecule gp82, fused to
GST. This T. cruzi protein, designated J18, which binds
to mammalian cells in a receptor-mediated manner, has
been shown to trigger Ca
2+
-dependent F-actin depo-
lymerization in epithelial HeLa cells [20]. We found out
that J18 induces actin cytoskeleton disruption in Tm5
but not in melan-a cells, leading to the apoptotic cell
death involving caspase-3 activity.
Materials and methods
Mice
Six-week-old female C57BL/6 mice from the animal
facility CEDEME, at Universidade Federal de Sa
˜
o Paulo,
were used for tumor-development experiments. All
procedures and experiments conformed to the regula-
tions of the institutional Ethical Committee for animal
experimentation.
Cell culture, proliferation and adhesion assays
The nontumorigenic murine melanocyte cell line melan-a
[18] was cultured in RPMI 1640 medium (Invitrogen,
Carlsbad, California, USA), pH 6.9, supplemented with
5% fetal calf serum and 200 nmol/l of 4a-phorbol
myristate 13-acetate (Sigma, St Louis, Missouri, USA)
at 371C in a humidified 5% of CO
2
atmosphere. The
anoikis-resistant tumor cell line Tm5, derived from
melan-a cells and selected upon apoptosis of melan-a
cells by adhesion blockade, as reported in an earlier study
[19], was cultured in the same conditions, without
4a-phorbol myristate 13-acetate. For cell proliferation
studies, 2.5 Â 10
3
cells were grown at 371C in 96-well
microtiter plates for different periods of time, depending
on the experiment. The number of viable cells was
estimated using a standard methyl thiazol tetrazolium
(MTT) assay (Sigma) and by reading the optical density
at 570 nm, after solubilization of formazan product with
isopropanol [21]. Adhesion assays were performed by
adding melan-a or Tm5 cells (1 Â10
5
per well) to 96-well
plates. After 20 min of incubation at 371C, washings with
phosphate-buffered saline (PBS) to remove nonadherent
cells, the viable adherent cells were quantified by MTT
assay.
Purification of J18 and gluthatione-S-transferase
The recombinant protein J18, containing the full-length
T. cruzi gp82 sequence (GenBank access number L14824)
in frame with GST, was produced in Escherichia coli (DH5-
a) by transforming the bacteria with a pGEX-3 construct
comprising the gp82 gene [22]. Purification of J18 protein
was performed as previously described [20]. To ascertain
that the correct protein was obtained, the purified sample
was analyzed by silver staining of sodium dodecyl sulfate–
polyacrylamide gel electrophoresis gel, as well as by
immunoblotting using monoclonal antibody 3F6 directed
to gp82. The same procedure to purify J18 was used to
obtain GST.
Preparation of antibodies to J18
To obtain antibodies to the recombinant protein J18,
BALB/c mice were injected intraperitoneally with the
protein (5 mg/mouse) adsorbed in aluminum hydroxide
as adjuvant. Two weeks after the first dose, the animals
received three additional weekly doses of the protein
(5 mg/mouse) in the same adjuvant. The mice were bled
by heart puncture 10 days after the last dose, the sera
were collected and stored at 201C until used.
Binding of J18 to cells and actin filaments visualization
For binding assay, melan-a or Tm5 cells (3 Â 10
4
cells per
well) were grown overnight in 96-well plates. After
fixation with 3.5% formaldehyde in PBS, blocking with
PBS containing 10% fetal calf serum, and incubation with
J18 or GST, for 1 h at 371C, the cells were sequentially
incubated with polyclonal antiserum directed to J18 or to
GST, and with peroxidase-conjugated anti-mouse IgG.
The bound enzyme was revealed using o-phenylene-
diamine as detailed elsewhere [23]. For actin filament
visualization, melan-a or Tm5 cells (5 Â 10
4
), grown in
13-mm diameter round glass coverslips placed in 24-well
plates, were incubated in absence or in the presence of
J18 or GST for 8 h. After fixation with 3.5% formaldehyde
in PBS, the cells were incubated for 1 h at room
temperature with phalloidin-rodhamin (Sigma) in PBS
containing 0.15% gelatin and 0.1% saponin, before
microscopic visualization. Images of stress fibers were
taken with a confocal system (Zeiss Axiovert 100 mol/l),
63 Â1.3 oil objective, using the software LSM 510 Expert
Mode SP2 (Carl Zeiss MicroImaging Inc., Thornwood,
New York, USA).
Determination of caspase-3 activity
Melan-a or Tm5 cells (1 Â 10
6
cells/well), grown over-
night at 371C in six-well plates, were treated or not with
0.8 mmol/l J18 or GST, at 371C for 36 h, and then
trypsinized for detachment from the substrate. The
activity of caspase-3 in whole cell lysates (2 Â 10
6
cells/
sample) was determined using colorimetric assay kits
(Calbiochem, UK), according to the manufacturer’s
instructions. Briefly, cells were lysed and centrifuged,
at 17 400g for 10 min at 41C, to remove the cell debris.
Trypanosoma cruzi surface molecule gp82 Atayde et al. 173
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
The supernatants were incubated for 6 h with 0.2 mmol/l
substrate for caspase-3 (p-nitroanilide-conjugated DEVD
peptide). Human recombinant caspase-3 served as
positive control and cell extracts treated with 0.1 mmol/l
of inhibitor for caspase-3 (DEVD-CHO) served as
negative control. The absorbance was measured at
405 nm using a 96-well colorimetric plate reader (Lab-
systems Multiskan MS, Helsinki, Finland).
Annexin-V assay for detection of phosphatidylserine,
tetramethyl-rhodamine ethyl ester assay for detection
of mitochondrial depolarization and cell-cycle analysis
Melan-a or Tm5 cells (1 Â 10
6
cells/well), grown in six-well
plates, were treated or not with 0.8 mmol/l of J18 or GST.
After 48 h or 72 h at 371C, depending on the experiment,
cells were trypsinized, collected, washed with PBS, and the
concentration adjusted to 1 Â 10
6
cells/ml. For phosphati-
dylserine (PS) detection, to 100 mlofcellsamples1mgof
annexin-V-FITC (Santa Cruz Biotechnology, California,
USA) and 10 mlof50mg/ml propidium iodide (PI; Sigma)
were added to 100 ml of cell samples. After incubation for
15 min at room temperature in the dark, flow cytometric
analysis was performed in a FACSCalibur flow cytometer
(Becton Dickinson Biosciences, Palo Alto, California, USA).
Data acquisition and analysis were performed by CellQuest
Pro software (BD Biosciences). The fluorescence intensity
of cells stained with annexin-V-FITC was detected in FL1
channel (525 nm), and PI was detected in the FL3 channel
(620 nm). To detect mitochondrial depolarization, the
mitochondria specific dye tetramethyl-rhodamine ethyl
ester (TMRE) (Molecular Probes-Invitrogen) was added
to 100 mlofcellsamples,kindlyprovidedbyDrSoraya
Smaili, at 100 nmol/l final concentration. After 20-min
incubation at 371C, flow cytometric analysis and data
acquisition were performed as above, by measuring the
fluorescenceofcellsstainedwithTMREinFL2channel
(575 nm). For cell-cycle analysis, cells were fixed with
ethanol (Merck, Darmstadt, Germany), washed with PBS
and treated with 100 mg/ml RNAse A (Sigma) in 100 mlPBS
for 5 min at room temperature. The cells were then
incubated with 50 mg/ml of PI for 30 min at 41C. Flow
cytometric analysis and data acquisition were performed as
in other two assays, by measuring the DNA content
of cells stained with PI in the FL2 channel (575 nm). The
number of apoptotic cells was determined by evaluating
the percentage of hypodiploid nuclei in the DNA less than
2N peak (sub-G1 population). In all three assays 10 000
events were collected from each sample.
Genomic DNA electrophoresis
Melan-a or Tm5 cells (1 Â 10
6
cells/well), grown in six-
well plates were treated or not with 0.8 mmol/l of J18 or
GST, for 72 h at 371C, before trypsinization. After washes
in PBS, cells were resuspended in TELT buffer (50 nmol/l
Tris, pH 8.0; 62.5 nmol/l EDTA, pH 9.0; 2.5 mol/l LiCl;
4% Triton X-100) and treated with 50 mg of RNAse A
(Sigma) for 1 h at 371C. After addition of 150 mlof
phenol : chloroform : isoamyl alcohol (USB Co, Cleveland,
Ohio, USA), the samples were centrifuged at 17 400g for
15 min. The aqueous phase, containing genomic DNA, was
transferred to new tube and 300 ml of absolute ethanol
(Merck) was added to each sample. After the DNA
precipitation for 20 min at 701C, the samples were
centrifuged at 17 400g for15min,washedin70%ethanol,
dried, and resuspended in TE buffer (10 mmol/l Tris,
1 mmol/l EDTA, pH 8.0). The extracted DNA was applied
in a 0.8% agarose gel, subjected to electrophoresis and the
ethidium bromide bands were visualized under UV light.
Visualization of the apoptotic cell nucleus and cellular
localization of NF-jB
For both assays, melan-a or Tm5 cells (5 Â 10
4
), grown in
13-mm diameter round glass coverslips in 24-well plates,
were treated or not with 0.8 mmol/l of J18 or GST. After
48 h or 72 h, depending on the experiment, the cells
were fixed with 3.5% formaldehyde in PBS for 1 h at
room temperature with 4
0
,6-diamidino-2-phenylindole
(200 mmol/l) in PBS containing 0.15% gelatin with 0.1%
saponin. Images of nuclei were taken at fluorescence
microscope using ACT-1 software. To detect the localiza-
tion of NF-kB, the cells were incubated with anti-NF-kB
monoclonal antibody (1 : 10) (Santa Cruz Biotechnology)
in the PBS containing 0.15% gelatin with 0.1% saponin,
followed by incubation with FITC-conjugated anti-mouse
IgG (Sigma) for 1 h at room temperature, before micro-
scopic visualization and capturing of images as above.
In-vivo tumor growth assay
Tm5 cells were harvested from trypsin-treated subcon-
fluent monolayers, counted, resuspended in PBS and 2 Â
10
5
cells were subcutaneously injected into the dorsal
area of 6–8-week-old female C57BL/6 mice, which were
then divided into four groups (n = 8 per group). Starting
on the day of tumor cell inoculation, each group received
either PBS, GST or J18, for 10 consecutive days, at the
site of Tm5 cell injection, except one group that received
J18 on the opposite side. The dose of J18 or GST was
1 mg/kg/day. Tumor size was monitored everyday and
determined as follows: [maximum diameter  (minimum
diameter)
2
]/2. Mice with a tumor mass larger than
10 mm
3
were considered positive for the presence of tumor.
Trypanosoma cruzi strains
T. cruzi strains G and Y* were used to infect mice. Y*
strains are a gp82-deficient variant of Y strain. The
parasites were maintained cyclically in mice and in liver
infusion tryptose medium containing 5% fetal bovine
serum. Metacyclic forms from cultures at the stationary
growth phase were purified by passage through DEAE-
cellulose column, before inoculation into mice.
Statistical analysis
For in-vivo experiments, statistical analysis was per-
formed using the two-way analysis of variance test
17 4 Melanoma Research 2008, Vol 18 No 3
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
followed by Bonferroni post-tests in the GraphPad Prizm
5.0 Software. To analyze the results of in-vitro experi-
ments, the program GraphPad InStat tm was used to
determine significance by Student’s t-test.
Results
The recombinant protein J18 inhibits the growth
of Tm5 melanoma cells
Melan-a melanocytes and melanoma lineage Tm5 were
treated for 48 h with varying concentrations of J18, which
contains the T. cruzi gp82 sequence fused to GST, and
the cell viability was determined by MTT assay. J18
displayed a significant inhibitory effect on the prolifera-
tion of Tm5 cells, but not of nontumorigenic melan-a
cells, at all concentrations tested (Fig. 1a). At 0.8 mmol/l,
the inhibition by J18 was higher than 60%, and this
concentration was used in all subsequent experiments.
GST was devoid of inhibitory effect. As the growth curves
show (Fig. 1b), the inhibition of Tm5 cell proliferation by
J18 became evident from 48 h postincubation, and by 72 h
the cells were dead.
When visualized using the contrast phase microscope,
J18-treated Tm5 cells, but not melan-a cells, displayed an
altered morphology, with cytoplasm shrinkage and
detachment from neighboring cells as well as from the
substrate (Fig. 1c). The effect of J18 on human tumor
cell lines was also tested. Proliferation of human
melanoma Mel85 cells, as well as of mammary adeno-
carcinoma MCF7 cells, was inhibited more than 60% after
72 h of incubation with J18. Growth inhibition of normal
mammary epithelial MCF10 cells by J18 was negligible.
Treatment of Tm5 melanoma cells with J18 leads to
disassembly of actin fibers and reduced substrate
adhesion
In cell binding assays, J18 bound to melan-a or Tm5 cells
in a dose-dependent manner (Fig. 2a), indicating that the
lack of growth inhibitory effect on melan-a cells (Fig. 1a
and b) is not the result of the differential binding. It has
previously been shown that J18 induces disruption
of actin cytoskeleton in HeLa cells [20]. To determine
whether such an effect was observed in Tm5 cells, the
architecture of actin cytoskeleton was visualized at
different time points after incubation with 0.8 mmol/l
J18, by staining with phalloidin-rhodamine. Within
30 min, F-actin rearrangement was detectable in Tm5
cells, which displayed diminished number of stress fibers,
the alterations becoming more pronounced as the
incubation time increased, whereas GST displayed no
such effect (Fig. 2b). Melan-a cells remained unaltered
by incubation with J18 (Fig. 2b). Treatment of cells with
J18 for 15 min before seeding into microtiter plates,
followed by 20 min of incubation to allow cell attachment,
resulted in diminished adhesion of Tm5 but not of
melan-a cells (Fig. 2c).
J18 induces apoptotic cell death of Tm5 melanoma cells
The F-actin-disrupting effect of J18 observed in HeLa
cells, the epithelial derived from human carcinoma, albeit
to a lesser degree, is similar to that of cytochalasin D [20].
As apoptosis induced by cytochalasin D has been reported
to occur in adherent epithelial cells [6] and also in human
megakaryoblastic leukemia CMK cells [7], we examined
whether the death of Tm5 cells treated with J18 occurred
by apoptotic process. We performed the assay using
FITC-labeled Annexin-V, a Ca
2+
-dependent phospho-
lipid-binding protein with high affinity for PS, which is
located in the inner side of the plasma membrane and
translocates to the external surface in the early stages of
apoptosis [24]. For cell staining, PI with red fluorescence
Fig. 1
0
50
Viable cells (OD
570
)
0.80.40.20.10
J18Control(c)
J18 (μmol/l) Time (h)
Tm5 Melan-a Viable cells (%)
100
Tm5
0
1
2
967248240
J18
GST
Control
(b)(a)
Melan-a
150
Growth inhibitory effect of J18 on Tm5 melanoma cells. (a) Melan-a or
Tm5 cells were incubated with different concentrations of J18 for 48 h
and viable cells were quantified by methyl thiazol tetrazolium assay. Tm5
growth inhibition was significant (*) at all J18 concentrations (P < 0.005
at 0.1 mmol/l and P < 0.0001 at 0.8 m mol/l). The reference value of
100% was ascribed to the untreated control. Values are mean ± SD of
three experiments performed in triplicates. (b) Cells were incubated or
not with J18 or glutathione-S-transferase (GST), at 0.8 mmol/l, for
different periods of time. Inhibition of Tm5 cell proliferation was highly
significant at 72 h and 96 h (P < 0.0001). Open and closed symbols
correspond respectively to melan-a and Tm5 growth curves.
Representative results of five experiments, expressed as the mean ± SD
of triplicates, are shown. (c) Phase contrast images showing
morphological changes in Tm5 cells treated or not with 0.8 mmol/l J18
for 72 h. Note the cytoplasm shrinkage and the detachment from the
neighboring cells. No alterations were seen in J18-treated melan-a
cells. Images are representative of four independent assays.
Trypanosoma cruzi surface molecule gp82 Atayde et al. 175
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
was used simultaneously, to discriminate apoptotic cells
from necrotic cells. Analysis by flow cytometry showed an
increased number of FITC
+
and PI-negative Tm5 cells
after treatment with J18 for 48 h (Fig. 3), indicating that
PS flipped without membrane damage. To a lesser
degree, there was also an increase in double positive
J18-treated Tm5, which probably represent a minor cell
population at more advanced apoptotic process, in which
the integrity of the membrane was lost. PS exposure was
not observed in melan-a cells treated with J18, or in cells
incubated with GST. Other changes associated with
apoptosis were also found specifically in J18-treated Tm5
cells. An increase in sub-G1 fraction (Fig. 4a), which
is associated with DNA fragmentation (Fig. 4b), was
detectable after 72 h of incubation of Tm5 cells with
J18. No such alteration was detected in melan-a cells
treated with J18 (Fig. 4a and b). The nuclei of
J18-treated Tm5 cells, examined by staining with DAPI,
exhibited altered morphology, aggregated chromatin and
blebs (Fig. 4c).
Fig. 2
Melan-a
J18
1.0
(c)(a)
0.5
0.0
Control GST J18
GST
Tm5
Melan-a
Tm5
1.2
0.6
Cell binding (OD
492
)Melan-aTm5
Substrate adhesion (OD
570
)
0.0
0 0.1 0.2
Protein (μmol/l)
Control(b) GST J18
0.4 0.8
J18-induced disassembly of actin fibers in Tm5 melanoma cells and reduced adhesion to substrate. (a) Melan-a or Tm5 cells were fixed and
incubated with varying concentrations of J18 or glutathione-S-transferase (GST). The binding was revealed by sequential incubation with polyclonal
antiserum directed to J18 or to GST and with peroxidase-conjugated anti-mouse IgG. Note the similar binding profile of J18 to both cells. The
representative results of three independent assays performed in triplicate are shown. Values are the mean ± SD of triplicates. (b) Confocal images
showing the phalloidin-rodhamin-stained actin filaments of melan-a or Tm5 cells incubated or not for 8 h with 0.8 mmol/l of J18 or GST. Note the
stress fibers decrease in Tm5 cells treated with J18, as compared with controls or to melan-a cells. Scale bar: 10 mm. Images are representative of
four independent experiments. (c) Substrate adhesiveness of melan-a or Tm5 cells treated or not for 15 min with 0.8 mmol/l of J18 or GST. Pretreated
cells were added to microtiter plates and incubated for 20 min at 371C for attachment. After removing nonadherent cells, adherent cells were
quantified by methyl thiazol tetrazolium assay. The differential adhesiveness of Tm5 cells treated with J18 was significant (P < 0.02). Data are the
mean ± SD of triplicates and representative of three independent experiments.
176 Melanoma Research 2008, Vol 18 No 3
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
Caspase-3 activity is increased in J18-treated Tm5
melanoma cells
We checked the involvement of caspases in the apoptotic
cell death of Tm5 cells, in particular of the effector
caspase-3, which has its activity increased upon treat-
ment of transformed cells with cytochalasin D [7].
Melan-a or Tm5 cells were incubated for 36 h with J18
or GST, and the activity of caspase-3 was measured.
In J18-treated Tm5 cells, but not in melan-a cells, there
was an increase in caspase-3 activity, which was
inhibitable by a specific inhibitor (Fig. 5). Provided that
mitochondria regulate the final execution of the apoptotic
program by releasing apoptogenic factors [25–27], and
cytochrome c-related caspase-3 activation have been
reported [28], we investigated the involvement of these
organelles at late stages of the Tm5 melanoma apoptotic
Fig. 3
Annexin-V
Tm5
ControlGSTJ18
Propidium iodide
8.00% 7.47%
0.49%
7. 4 8 %
54.69%
19.39%2.23%8.21%2.98%
0.51%
1.41%
9.62%8.31%2.47%5.60%
0.36%
1.44%
8.77%
Melan-a
10
4
10
3
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
10
4
10
3
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
10
4
10
3
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
10
4
10
3
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
10
4
10
3
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
10
4
10
3
10
2
10
1
10
0
10
0
10
1
10
2
10
3
10
4
Phosphatidylserine exposure on Tm5 melanoma cells treated with J18. Melan-a or Tm5 cells were treated or not with 0.8 mmol/l of J18 or glutathione-
S-transferase (GST) for 48 h. Cells were collected and incubated with annexin-V-FITC and propidium iodide (PI) for 15 min at room temperature in
the dark. Flow cytometric analysis was performed on a FACSCalibur flow cytometer. The fluorescence of cells stained with annexin-V-FITC was
detected in FL1 channel (525 nm), whereas PI was detected in the FL3 channel (620 nm). Ten thousand events were collected from each sample.
Graphics shown are representative of three independent experiments.
Trypanosoma cruzi surface molecule gp82 Atayde et al. 177
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
pathway. By using the fluorescent probe TMRE, which
accumulates rapidly and reversibly in mitochondria [29],
we examined the loss of mitochondria transmembrane
potential, which is one of the events in apoptosis [30].
In Tm5 cells treated with J18 for 72 h, but not in
J18-treated melan-a cells or GST-treated cells, mito-
chondrial depolarization was detected to some extent
(Fig. 6).
Translocation of NF-jB to the nucleus is reduced
in Tm5 melanoma cells treated with J18
Constitutive activation of the nuclear transcription factor
NF-kB, which is normally sequestered in the cytosol and
translocates to the nucleus upon degradation of its
inhibitor IkB, is an emerging hallmark of various types
of tumor, including melanoma [31]. Nuclear localization
and NF-kB-binding to DNA is enhanced in melanoma
Fig. 4
10008006004002000
0
20
40
60
80
100
10008006004002000
0
20
40
60
80
100
10008006004002000
0
20
40
60
80
100
Tm5 J18
Tm5 GST
Tm5 control
M-a J18
M-a GST
M-a control
Propidium iodide
1000
3
4
2
1
8006004002000
0
20
0.1
0.4
2.0
4.3
6.5
23.1
40
60
80
100
10008006004002000
0
20
40
60
Cell counts
80
100
10008006004002000
0
20
40
60
80
3.6%
0.8%
62.1%
18.5%
2.6%
J18 GST Control
Tm5
(c)(b)
(a) Melan-a
0.7%
100
Increase in sub-G1 population in Tm5 melanoma cells treated with J18. (a) Cell-cycle analysis of melan-a or Tm5 cells treated or not with 0.8 mmol/l of
J18 or glutathione-S-transferase (GST) for 72 h. Fixed cells were stained with propidium iodide and analyzed by flow cytometry. The DNA content of
cells was detected in the FL2 channel (575 nm). The number of DNA-fragmented cells was determined by evaluating the percentage of less than 2N
DNA peak (sub-G1 population). Ten thousand events were collected from each sample. The representative of three independent experiments are
shown. (b) Genomic DNA extracted from melan-a or Tm5 cells treated or not with 0.8 mmol/l of J18 or GST for 72 h was applied in a 1.0% agarose
gel and the ethidium bromide bands were seen under UV light. The same profile was seen in two other experiments. (c) Visualization of apoptotic
nuclei of Tm5 cells treated with 0.8 mmol/l J18 for 72 h. Fixed cells were stained with 4
0
,6-diamidino-2-phenylindole and images were obtained at
fluorescence microscope utilizing the ACT-1 software. Representative images: (1: kidney shaped; 2: chromatin aggregation; 3: nuclear blebbing; 4:
nuclear blebs) of three independent experiments are shown.
178 Melanoma Research 2008, Vol 18 No 3
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
cells, as compared with normal melanocytes [32]. Here,
we examined whether treatment of Tm5 cells with
J18 interfered with nuclear localization of NF-kB. Melan-
a and Tm5 cells were incubated with J18 for 48 h and
then processed for detection of NF-kB using monoclonal
antibody directed to NF-kB subunit p50. The images
visualized using the fluorescence microscope showed a
predominant cytoplasmic localization of NF-kB in J18-
treated Tm5 cells at 48 h postincubation, whereas much
fewer J18-treated melan-a cells were negative for nuclear
NF-kB (Fig. 7).
J18 reduces the tumor growth in mice inoculated with
Tm5 melanoma cells
Experiments were also performed to test the antitumor
activity of J18 in vivo. C57BL/6 mice were injected
subcutaneously with Tm5 cells in the dorsal area and
then separated in four groups. For 10 consecutive days,
starting on the day of Tm5 cell inoculation, one group of
mice was treated with PBS, the second group with GST,
the third group with J18, all of them at the site of
Fig. 6
100
80
12.1%
Melan-a Tm5
21.3%
23.8%
60
40
20
0
10
0
10
1
10
2
10
3
10
4
100
80
60
40
20
0
10
0
10
1
10
2
10
3
10
4
100
80
Cell counts
ControlGSTJ18
15.1%
60
40
20
0
10
0
10
1
10
2
10
3
10
4
100
80
60
40
20
0
10
0
10
1
10
2
10
3
10
4
100
80
17.2%
60
40
20
0
10
0
10
1
10
2
10
3
10
4
TMRE
100
80
38.8%
60
40
20
0
10
0
10
1
10
2
10
3
10
4
Reduction of mitochondrial polarity in Tm5 melanoma cells treated with J18. Melan-a or Tm5 cells, treated or not with 0.8 mmol/l of J18 or glutathione-
S-transferase (GST) for 72 h, were collected and incubated with tetramethyl-rhodamine ethyl ester (TMRE) for 20 min at 371C. Flow cytometric
analysis was performed in a FACSCalibur flow cytometer. The fluorescence of cells stained with TMRE was detected in FL2 channel (575 nm).
Ten thousand events were collected from each sample. Shown are the representative results of three independent experiments.
Fig. 5
0.4
0.2
Caspase–3 activity (OD
405
)
0.0
Melan-a Tm5
Control
GST
J18
J18+C3-I
Caspase-3 activation in Tm5 melanoma cells treated with J18. The
activity of caspase-3 in whole cell lysates was determined using a
colorimetric assay. Cells incubated or not for 36 h with 0.8 mmol/l J18
or glutathione-S-transferase (GST) were lysed. Supernatants were then
incubated with caspase-3 substrate (DEVD-p-nitroanilide) or with
caspase-3 substrate in the presence of caspase-3 inhibitor (C3-I/
DEVD-CHO), for 6 h at 371C. The absorbance of each sample was
measured at 405 nm. Data are the means of triplicates from one out
of three independent experiments.
Trypanosoma cruzi surface molecule gp82 Atayde et al. 179
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
Fig. 7
Control
NF-κBNF-κBDAPI DAPI
Melan-aTm5
J18
NF-kB localization in Tm5 melanoma cells treated with J18. Melan-a or Tm5 cells were incubated or not for 48 h with 0.8 mmol/l of J18 or glutathione-
S-transferase (GST), fixed and incubated sequentially with anti-NF-kB monoclonal antibody, FITC-conjugated anti-mouse IgG and 4
0
,6-diamidino-2-
phenylindole (DAPI). Note the predominant cytoplasmic localization of NF- kB in Tm5 cells treated with J18. No alterations were seen in cells treated
with GST. Images, obtained with the ACT-1 software, are representative of three independent experiments. White and black arrows point to nuclei
negative and weakly positive for NF-kB, respectively.
Fig. 8
6000
Control
GST
J18 control
J18
Tumor volume (mm
3
)
Tumor volume (mm
3
)
5000
4000
3000
2000
1000
0
Time (days)
10
5000
(b)
(a)
100
75
50
25
0
0102030
Time (days)
40 50
PBS
J18 (8 23)
J18 (1 10)
4000
3000
2000
1000
9111315
Time (days)
17 19 21 23
0
12 14 16 18 20
100
80
Percent survival
Percent survival
60
40
20
0
01020
Time (days)
30 40
Reduction of in-vivo tumor growth in mice treated with J18. (a) Tm5 cells (2 Â 10
5
) were injected subcutaneously into the dorsal area of C57BL/6
mice ( n = 8 per group). The mice were divided into four groups: phosphate-buffered saline (PBS) treated, glutathione-S-transferase (GST) treated,
J18 treated outside the tumor area (J18 control), and J18 treated in tumor area (J18). Treatment with J18 or GST (1 mg/kg/day) was performed for 10
consecutive days beginning on the day of melanoma cell inoculation. Development of tumor is shown on the left and mortality on the right. Data are
the mean ± SD of tumor volumes of eight animals from each group, and are representative of three independent experiments. The tumor growth
inhibition in J18 group from day 17 to 20 was significant (P < 0.001). (b) C57BL/6 mice were injected subcutaneously with Tm5 cells (2 Â 10
5
) and
separated in three groups (n = 5 per group): PBS treated, J18 treated in tumor area starting on day zero of melanoma cell inoculation, and J18
treated from day 8 to 23. Data are the mean ± SD of tumor volumes of five animals from each group. Inhibition of tumor growth in J18 group was
significant from day 21 to 23 (P < 0.001).
180 Melanoma Research 2008, Vol 18 No 3
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
melanoma cell injection, whereas the fourth group
received J18 on the opposite side of Tm5 cell inoculation.
The daily dose of either of J18 or GST was 1 mg/kg/day.
Of the four groups, only the mice inoculated with J18
proximal to the site of melanoma cell injection developed
tumors of smaller size than the controls and survived
longer (Fig. 8a). We confirmed that J18, by itself, is not
harmful to the animals. When injected daily with J18,
following the same procedure for the experiments
described above, the mice remained healthy during a
period of 60 days, without signs of harmful effects, as if
they had not received any treatment. To test the effect of
J18 in vivo on more established melanoma, a group of mice
was inoculated with tumorigenic Tm5 cells and subjected
to treatment with J18 starting on day 8 postinoculation.
This scheme was as effective in delaying tumor develop-
ment as the protocol of treatment with J18 starting on the
day of Tm5 cell inoculation (Fig. 8b).
Tumor development is slower in mice infected with
T. cruzi strain expressing high levels of gp82 than in
mice infected with gp82-deficient strain
In addition to the effect of J18, we also investigated
whether mice inoculated with metacyclic forms of T. cruzi
strains expressing or deficient in gp82, the surface
molecule on which J18 is based, displayed differential
tumor development after injection of Tm5 cells. C57BL/
6 mice were separated in three groups. One of them was
inoculated intraperitoneally with metacyclic forms (10
6
/
mouse) of T. cruzi strain G, which express high gp82 levels
on the surface [33], whereas the other group received
metacyclic forms of strain Y*, which are deficient in gp82
(Fig. 9a). Three weeks postinfection, all groups were
injected subcutaneously with Tm5 cells in the dorsal
area. Tumor development was slower in mice infected
with G strain as compared with mice infected with Y*
strain or uninfected controls (Fig. 9b).
Discussion
Our study shows that a recombinant protein, J18, based
on the T. cruzi surface molecule gp82, is capable of
specifically inducing the apoptotic cell death of melano-
ma cells without exerting any effect on normal melano-
cytes. Since the early findings, more than 70 years ago,
that T. cruzi infection conferred resistance to tumor
development in mice, studies using killed parasites or
their extracts either in cancerous animals or in-vitro
cancer cell cultures have been performed, with con-
troversial results [12,13,34]. Although the possibility of T.
cruzi factors with direct toxic activity toward tumor cells
has been raised, investigations on the antitumor proper-
ties of molecularly defined T. cruzi components are just
beginning. Recently, a T. cruzi protein Tc52 fused to GST
was shown to exhibit an apoptosis-inducing effect on a
human T-cell leukemic cell line [14]. Other T. cruzi
molecules, such as the ceramide-containing glycolipid
and trans-sialidase were reported to have proapoptotic
activity toward cells of the immune system [15–17]. It
remains to be determined whether they possess anti-
tumor activity.
What is relevant about the effect of J18, the protein based
on T. cruzi surface molecule gp82, which also acts in vivo,
is its capacity of inhibiting the growth and ultimately
leading to death the tumorigenic Tm5 cells, but not the
melan-a cells, the normal melanocytes from which they
derived. J18 induced actin cytoskeleton disruption in
Tm5 cells and this is possibly associated with its
apoptosis-promoting activity. Apoptosis of different cell
types, under various conditions, has been shown to occur
after treatment with cytochalasin D or jasplakinolide,
drugs that affect the cellular actin architecture [6–9].
Several evidences indicate that J18 induces apoptosis in
Tm5 cells. Flipping of phosphatidylserine from the inner
to the external side of the plasma membrane was
detected in J18-treated Tm5 melanoma cells but not in
melan-a melanocytes. A number of other events asso-
ciated with the apoptotic process, such as alteration in
nuclear morphology, DNA fragmentation, and decrease in
mitochondria transmembrane potential, were also found
to occur specifically in J18-treated Tm5 cells. The
apoptotic pathway triggered by J18 possibly involves the
executioner caspase-3. In J18-treated Tm5 cells, the
activity of caspase-3 was increased. Another finding that
reinforces the apoptosis-inducing activity of J18 toward
Tm5 cells is that NF-kB is mostly retained in the
Fig. 9
T. cruzi
G
gp82
(a) (b)
gp30
Y
2000
Control
G strain
Y
strain
1500
1000
Tumor volume (mm
3
)
500
0
91113
Time (days)
15 17 19
Tumor development in mice infected with gp82-expressing or gp82-
deficient T. cruzi strains. (a) Metacyclic trypomastigotes of G and Y*
strains were processed for western blot analysis using monoclonal
antibody 3F6 directed to the surface molecule gp82. Note the
expression of high levels of gp82 in G strain and the detection of gp30,
instead of gp82, in Y* strain. (b) Groups of C57BL/6 mice were
separated in three groups (n = 5 per group) and were inoculated with
metacyclic forms (10
6
/mouse) of T. cruzi strain G or Y*, or remained
uninfected. Three weeks later, all animals were injected subcutaneously
with Tm5 cells (2 Â 10
5
). Data are the mean ± SD of tumor volumes of
five animals from each group. The delay in tumor development in mice
infected with T. cruzi G strain was significant from day 15 to 18
(P < 0.001).
Trypanosoma cruzi surface molecule gp82 Atayde et al. 181
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
cytoplasm rather than translocating into the nucleus. The
activation of this nuclear transcription factor is associated
with the growth of various types of tumor, including
melanoma [31,32]. NF-kB activity was found to be
upregulated in lesions of melanoma when compared with
melanocytes in normal skin [32,35,36]. Taken together,
our data suggest that J18 specifically exerts its activity on
Tm5 melanoma cells not only by inducing the apoptotic
process, but and also by inhibiting cell growth through its
negative effect on NF-kB translocation to the nucleus.
Another interesting finding is that in mice infected with
metacyclic trypomastigotes of T. cruzi strain G, but not
of strain Y*, before injection of Tm5 melanoma cells,
a significant delay of tumor development was observed.
The main difference in the surface profile of the infective
metacyclic forms of strains G and Y* is the expression of
gp82, the surface molecule defined by monoclonal
antibody 3F6 [37] on which J18 protein is based. Instead
of gp82, Y* metacyclic forms express the surface
glycoprotein gp30 that also reacts with monoclonal
antibody 3F6 and, as gp82, is involved in host cell
invasion [33]. Gp82 is a member of a large gp85/sialidase
multigene family [38]. Included in this superfamiIy are
Tc-85 and trans-sialidase (TS). In this regard, it is of
interest that TS, similarly to gp82, was shown to have
proapoptotic activity [16,17]. TS also known as SAPA,
which is linked to the outer leaflet of the parasite plasma
membrane through glycophosphatidylinositol anchor
[39], is secreted and can be found in the blood of
T. cruzi-infected patients at the acute phase [40]. Tc85,
a surface glycoprotein expressed in parasite forms
corresponding to the bloodstream trypomastigotes, con-
tains binding sites for laminin [41,42] and is implicated in
mammalian cell entry [43]. To what extent Tc85 exhibits
the signaling properties similar to those displayed by
metacyclic trypomastigote gp82, and whether it could
inhibit tumor development, has not been investigated.
The scenario we envisage when mice infected with
T. cruzi are inoculated with cancer cells is that molecules
of gp85/sialidase family, which are preferentially ex-
pressed in infective trypomastigote forms that circulate
in the blood, act on tumor cells, interfering with their
development through growth inhibition and apoptotic
mechanisms, such as described for J18. Another possibi-
lity is that the immune response elicited by T. cruzi is also
effective toward tumor cells. A third possibility is that
both the toxic activity of T. cruzi factors and the parasite-
induced immune effectors are acting in concert.
Acknowledgements
This work was supported by Fundac¸a
˜
o de Amparo a
`
Pesquisa do Estado de Sa
˜
o Paulo (FAPESP) and Conselho
Nacional de Desenvolvimento Cientı
´
fico e Tecnolo
´
gico
(CNPq), Brasil. The authors thank Dr Erika Suzuki for
critical reading of the manuscript.
References
1Flo
´
rez A, Cruces M. Melanoma epidemic: true or false? Int J Dermatol 2004;
43:405–407.
2 Lens MB, Dawes M. Global perspectives of contemporary epidemiologica l
trends of cutaneous malignant melanoma. Br J Dermatol 2004; 150:
179–185.
3 Hanahan D, Weinberg RA. The hallmarks of cancer. Cell 2000; 100:57–70.
4 Chang HY, Yang X. Proteases for cell suicide: functions and regulation of
caspases. Microbiol Mol Biol Rev 2000; 64:821–846.
5 Gourlay CW, Ayscough KR. The actin cytoskeleton: a key regulator of
apoptosis and ageing? Nat Rev Mol Cell Biol 2005; 6:583–589.
6 White SR, Williams P, Wojcik KR, Sun S, Hiemstra PS, Rabe KF, et al.
Initiation of apoptosis by actin cytoskeletal derangement in human airway
epithelial cells. Am J Respir Cell Mol Biol 2001; 24:282–294.
7 Yamazaki Y, Tsugura M, Zhou D, Fujita Y, Shang X, Dang Y, et al. Cytoskeletal
disruption accelerates caspase-3 activation and alters the intracellular
membrane reorganization in DNA damage-induced apoptosis. Exp Cell Res
2000; 259:64–78.
8 Posey SC, Bierer BE. Actin stabilization by jasplakinolide enha nces
apoptosis induced by cytokine deprivation. J Biol Chem 1999;
274:4259–4265.
9 Odaka C, Sanders ML, Crews P. Jasplakinolide induces apoptosis in various
transformed cell lines by a caspase-3-like protease-dependen t pathway. Clin
Diagn Lab Immunol 2000; 7:947–952.
10 Hunter CA, Yu D, Gee M, Ngo CV, Sevignani C, Goldschmidt M, et al.
Cutting edge: systemic inhibition of angiogenesis underlies resistance to
tumors during acute toxoplasmosis. J Immunol 2001; 166:5878–5881.
11 Oliveira EC, Leite MS, Miranda JA, Andrade AL, Garcia SB, Luquetti AO,
et al. Chronic Trypanosoma cruzi infection associated with low incidence of
1,2-dimethylhydrazine-induced colon cancer in rats. Carcinogenesis 2001;
22:737–740.
12 Sheklakova LA, Kallinikova VD, Karpenko LP. Genetic heterogeneity of
Trypanosoma cruzi and its direct anticancer effect in cultured human tumor
cells. Bull Exp Biol Med 2003; 135:89–92.
13 Kallinikova VD, Matekin PV, Ogloblina TA, Leikina MI, Kononenko AF,
Sokolova NM, et al. Anticancer properties of flagellate protozoan
Trypanosoma cruzi. Izv Akad Nauk Ser Biol 2001; 3:299–311.
14 Borges M, Cordeiro-da-Silva A, Vergnes B, Sereno D, Ouaissi A. Conversion
of Trypanoso ma cruzi Tc52 released factor to a protein inducing apoptosis.
Tissue Cell
2005; 37:469–478.
15 Freire-de-Lima CG, Nunes MP, Corte-Real S, Soares MP, Previato JO,
Mendonc¸a-Previato L, et al. Proapoptotic activity of a Trypanosoma cruzi
ceramide-containing glycolipid turned on in host macrophages by IFN-
gamma. J Immunol 1998; 161:4909–4916.
16 Leguizamon MS, Mocetti E, Garcia Rivello H, Arqibay P, Campetella O.
Trans-sialidase from Trypanosoma cruzi induces apoptosis in cells from the
immune system in vivo. JInfectDis1999; 180:1398–1402.
17 Mucci J, Risso MG, Leguizamon MS, Frasch AC, Campetella O. The trans-
sialidase from Trypanosoma cruzi triggers apoptosis by target cell sialylation.
Cell Microbiol 2006; 8:1086–1095.
18 Bennett DC, Cooper PJ, Hart IR. A line of non-tumorigenic mouse
melanocytes, syngeneic with the B16 melanoma and requiring a tumor
promoter for growth. Int J Cancer 1987; 39:414–418.
19 Oba-Shinjo SM, Correa M, Ricca TI, Molognoni F, Pinhal MA, Neves IA, et al.
Melanocyte transformation associated with substrate adhesion impediment.
Neoplasia 2006; 8:231–241.
20 Cortez M, Atayde VD, Yoshida N. Host cell invasion mediated by
Trypanosoma cruzi surface molecule gp82 is associated with F-actin
disassembly and is inhibited by enteroinvasive Escherichia coli. Microbes Inf
2006; 8:1502–1512.
21 Mosmann T. Rapid colorimetric assay for cellular growth and survival:
application to proliferation and cytotoxicity assays. J Immunol Methods
1983; 65:55–63.
22 Santori FR, Dorta ML, Juliano L, Juliano MA, Franco da Silveira J, Ruiz RC,
et al. Identification of a domain of Trypanosoma cruzi metacyclic
trypomastigote surface molecule gp82 required for attachment and
invasion of mammalian cells. Mol Biochem Parasitol 1996; 78:
209–216.
23 Ramirez MI, Ruiz RC, Araya JE, Franco da Silveira J, Yoshida N. Involvement
of the stage-spec ific 82-kDa adhesion molecule of Trypanosoma cruzi
metacyclic trypomastigotes in host cell invasion. Infect Immun 1993;
6:3636–3641.
24 Vermes I, Haanen C, Steffens-Nakken H, Reutelingsperger C. A novel assay
for apoptosis. Flow cytometric detection of phosphatidylserine expression
on early apoptotic cells using fluorescein labelled Annexin V. J Immunol
Methods 1995; 184:39–51.
182 Melanoma Research 2008, Vol 18 No 3
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
25 Liu X, Kim CN, Yang J, Jemmerson R, Wang X. Induction of apoptotic
program in cell-free extracts: requirement for dATP and cytochrome c. Cell
1996; 86:147–157.
26 Slee EA, Harte MT, Kluck RM, Wolf BB, Casiano CA, Newmeyer DD, et al.
Ordering the cytochrome c-initiated caspase cascade: hierarchical
activation of caspases-2, -3, -6, -7, -8, and -10 in a caspase-9-dependent
manner. J Cell Biol 1999; 144:281–292.
27 Kroemer G, Reed JC. Mitochondrial control of cell death. Nat Med 2000;
6:513–519.
28 Meyer LH, Karawajew L, Schrappe M, Ludwig WD, Debatin KM, Stahnke K.
Cytochrome c-related caspase-3 activation determines treatment response
and relapse in childhood precursor B-cell ALL. Blood 2006; 107:
4524–4531.
29 O’Reilly CM, Fogarty KE, Drummond RM, Tuft RA, Walsh JV. Quantitative
analysis of spontaneous mitochondrial depolarizations. Biophys J 2003;
85:3350–3357.
30 Green DR, Reed JC. Mitochondria and apoptosis. Science 1998;
281:1309–1312.
31 Amiri KI, Richmond A. Role of nuclear factor-kappa B in melanoma. Cancer
Metastasis Rev 2005; 24:301–313.
32 Dhawan P, Richmond A. A novel NF-kappa B-inducing kinase-MAPK
signaling pathway up-regulates NF-kappa B activity in melanoma cells.
J Biol Chem 2002; 277:7920–7928.
33 Yoshida N. Molecular basis of mammalian cell invasion by Trypanosoma
cruzi. Ann Braz Acad Sci 2006; 78:87–111.
34 Cabral HR. The tumoricidal effect of Tr ypan os oma c ru zi: its intracellular cycle
and the immune response of the host. Med Hypotheses 2000; 54:1–6.
35 McNulty SE, Tohidian NB, Meyskens FL Jr. RelA, p50 and inhibitor of
kappa B alpha are elevated in human metastatic melanoma cells and
respond aberrantly to ultraviolet light B. Pigment Cell Res 2001;
14:456–465.
36 McNulty SE, del Rosario R, Cen D, Meyskens FL Jr, Yang S. Comparative
expression of NF-kappaB proteins in melanocytes of normal skin vs. benign
intradermal naevus and human metastatic melanoma biopsies. Pigment Cell
Res 2004; 17:173–180.
37 Teixeira MMG, Yoshida N. Stage-specific surface antigens of metacyclic
trypomastigotes of Trypanosoma cruzi identified by monoclonal antibodies.
Mol Biochem Parasitol 1986; 18:271–282.
38 Araya JE, Cano MI, Yoshida N, Franco da Silveira J. Cloning and
characterization of a gene for stage specific 82-kDa surface antigen of
metacyclic trypomastigotes of Trypanosoma cruzi. Mol Biochem Parasitol
1994; 65:161–169.
39 Schenkman S, Eichinger D, Pereira MEA, Nussenzweig V. Structural and
functional properties of Trypanosoma trans-sialidase.
Annu Rev Microbiol
1994; 48:499–523.
40 Affranchino JL, Ibanez CF, Luquetti AO, Rassi A, Reyes MB. Identification
of a Trypanosoma cruzi antigen that is shed during the acute phase of
Chagas’ disease. Mol Biochem Parasitol 1989; 34:221–228.
41 Giordano R, Chammas R, Veiga SS, Colli W, Alves MJM. Trypanosoma cruzi
binds to laminin in a carbohydrate-independent way. Braz J Med Biol Res
1994; 27:2315–2318.
42 Giordano R, Fouts DL, Tewari D, Colli W, Manning JE, Alves MJM. Cloning
of a surface membrane glycoprotein specific for the infective form of
Trypanosoma cruzi having adhesive properties to laminin. J Biol Chem
1999; 274:3461–3468.
43 Magdesian MH, Giordano R, Ulrich H, Juliano MA, Juliano L, Schumacker RI,
et al. Infection by Trypanosoma cruzi: identification of a parasite ligand and
its host-cell receptor. J Biol Chem 2001; 276:19382–19389.
Trypanosoma cruzi surface molecule gp82 Atayde et al. 183
Copyright © Lippincott Williams & Wilkins. Unauthorized reproduction of this article is prohibited.
Livros Grátis
( http://www.livrosgratis.com.br )
Milhares de Livros para Download:
Baixar livros de Administração
Baixar livros de Agronomia
Baixar livros de Arquitetura
Baixar livros de Artes
Baixar livros de Astronomia
Baixar livros de Biologia Geral
Baixar livros de Ciência da Computação
Baixar livros de Ciência da Informação
Baixar livros de Ciência Política
Baixar livros de Ciências da Saúde
Baixar livros de Comunicação
Baixar livros do Conselho Nacional de Educação - CNE
Baixar livros de Defesa civil
Baixar livros de Direito
Baixar livros de Direitos humanos
Baixar livros de Economia
Baixar livros de Economia Doméstica
Baixar livros de Educação
Baixar livros de Educação - Trânsito
Baixar livros de Educação Física
Baixar livros de Engenharia Aeroespacial
Baixar livros de Farmácia
Baixar livros de Filosofia
Baixar livros de Física
Baixar livros de Geociências
Baixar livros de Geografia
Baixar livros de História
Baixar livros de Línguas
Baixar livros de Literatura
Baixar livros de Literatura de Cordel
Baixar livros de Literatura Infantil
Baixar livros de Matemática
Baixar livros de Medicina
Baixar livros de Medicina Veterinária
Baixar livros de Meio Ambiente
Baixar livros de Meteorologia
Baixar Monografias e TCC
Baixar livros Multidisciplinar
Baixar livros de Música
Baixar livros de Psicologia
Baixar livros de Química
Baixar livros de Saúde Coletiva
Baixar livros de Serviço Social
Baixar livros de Sociologia
Baixar livros de Teologia
Baixar livros de Trabalho
Baixar livros de Turismo