Download PDF
ads:
______________________________________________________________________
Marcio Vinicius Bertacine Dias
Estudo Estrutural de Proteínas Alvo
de Mycobacterium tuberculosis
Tese apresentada ao Departamento
de Física do Instituto de
Biociências, Letras e Ciências
Exatas (IBILCE) da Universidade
Estadual Paulista “Júlio de
Mesquita Filho” (UNESP) para
obtenção de título de Doutor em
Biofísica Molecular.
Orientador: Prof. Dr. João Rugguiero Neto
Co-orientador: Prof. Dr. Walter Filgueira de Azevedo Jr.
São José do Rio Preto – SP
2007
______________________________________________________________________
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
O presente trabalho foi realizado no Laboratório de Sistemas Biomoleculares (BMSys),
Departamento de Física, Instituto de Biociências, Letras e Ciências Exatas/IBILCE de
São José do Rio Preto, SP, Universidade Estadual Paulista/UNESP, sob orientação do
Prof. Dr. João Ruggiero Neto e co-orientação do Prof. Dr. Walter Filgueira de Azevedo
Júnior, com auxílio da Fundação de Amparo à Pesquisa do Estado de São
Paulo/FAPESP (processo 03/12472-2).
ads:
À minha amada e querida mãe, Catarina,
E em memória ao meu inestimável pai, David
,
Dedico este trabalho aos amigos
Aos que se tornaram familiares,
Aos que nasceram familiares
E aos que conheci antes de ontem.
Dedico tanto aos que me deixam louco,
Quanto aos que enlouqueço.
Aos que me criticam em tudo,
E a um ou outro que atura
Minha “chatura”.
Aos amigos que correm,
Aos amigos que contemplam.
Aos que me consideram muito,
E aos muitos que, com razão, fazem pouco.
Aos que conhecem o que penso,
E aos que só conhecem o que faço.
Aos que passam o dia todo comigo,
E aos que estão todo tempo em mim.
Este trabalho é a soma de todos vocês.
E se ele não é melhor,
É por falta de memória,
Mas não por falta de amigos.
(autor desconhecido)
Agradecimentos:
Ao final deste trabalho são inúmeras as pessoas que eu devo o meu sincero obrigado, mas
entre estas as que me veio à mente no momento que eu escrevi esta parte do trabalho, sem
serem mais ou menos importantes as que infelizmente, pelo meu egoísmo ou esquecimento,
deixei de colocar neste local, são:
Ao meu orientador Prof. Dr. João Ruggiero Neto, por ter me socorrido no momento que
mais precisei; Pela amizade e confiança depositada em mim;
Ao meu co-orientador, Prof. Dr. Walter F. de Azevedo Jr., pela amizade, orientação,
ensinamentos, oportunidades e confiança depositada em mim;
Ao Prof. Dr. Valmir Fadel, pela amizade, apoio, conselhos e confiança depositada em
mim e pelo auxílio nas coletas realizadas no LNLS;
À Fernanda Canduri, pela amizade, apoio, conselhos, sendo que mesmo distante sempre
esteve perto nos momentos que eu mais precisei;
Às minhas ex-orientadoras, Profas. Dras. Marlene K. H. Kobayashi, Hermione Elly
Melara de Campos Bicudo, Ana Elizabete Silva e Fernanda Canduri, por a cada uma de
sua maneira, por todos os conhecimentos transmitidos e por me despertarem o gosto pela
ciência;
Às minhas queridas “alunas”, Brianna, Lívia, Michele e Samantha, que estiveram
sempre comigo me ensinando a ensinar e sempre trabalhando com dedicação e afinco no
Laboratório. E que sem essas maravilhosas pessoas grande parte deste trabalho talvez
não teria sido efetuada;
Ao meu grande amigo José Henrique Pereira, por toda amizade, conselhos, e ajuda tanto
na vida profissional como pessoal. A sua esposa, Adriana e sua família, que se tornou
uma extensão da minha própria família;
Ao Prof. Dr. Jorge Chahine e a Profa. Dra. Fernanda Canduri, por terem participado da
minha banca de qualificação;
Aos Prof(a)s Dr(a)s. Beatriz Guimarães, João Alexandre Barbosa, Maria Célia
Bertolini e Valmir Fadel por terem aceitado o convite em participar da minha banca de
doutorado; E também aos Prof(a)s Dr(a)s Marcos Fontes, Maria Cristina Nonato e
Jorge Chahine por terem aceitado o convite de serem suplentes da minha banca
examinadora;
À Marisa, minha leal suportadora, que sempre me agüentou na sala de estudos. Pelos
momentos de risadas, tristezas e desabafos; Pela amizade e sinceridade que sempre teve
comigo;
Aos Profs. Drs. Diógenes Santiago Santos e Luiz Augusto Basso, pelo fornecimento de
proteínas;
Ao pessoal da PUC-Rio Grande do Sul, pela amizade e pela expressão das proteínas
fornecidas, em especial, Fernanda Ely, Igor, Jaím, Ana Luíza, Patrícia, e Jordana;
À grande amiga Márcia, que sempre esteve disposta a me ajudar e me escutar. E que fez
também o ambiente de trabalho ser mais gostoso e profissional;
À outra grande amiga Denise Mello, que também sempre fez a minha vida se tornar mais
alegre;
À Daiana e Angélica por todo o carinho, simpatia e confiança;
A todos os alunos do Grupo BMSys tanto os atuais como os que hoje não freqüentam
mais (Pepeu, Diego, Denis, Janaína, Nathália, Nelsinho, Nelsão, Henrique, Marisa,
Galo, Paulo, Alexandre, Hugo, Helen, Alessandra, Bona, Joane, Guilherme, Daiana,
Ana Helena, Karisa, José Renato, Cristiane, Mariane, Lisandra, Marcos Michel,
Brianna, Angélica, Lívia, Michele e Samantha), pelas gargalhadas, risadas, choros,
brigas e tudo mais que faz parte das novas vidas;
Aos grandes amigos da pós-graduação, por tornar a vida cotidiana menos monótona e
solitária (Priscila, Gisele, Magno, Flávio, Sabrina, Jorge, Eduardo, Sídney, Leandro,
Ronaldo, Ricardo, José Ésio, Aperreio, Cristina, Fernanda, André, Renato);
Ao Prof. Dr. Jorge Chahine, pelo empenho com a pós-graduação e pelo exemplo de
pessoa que é;
Aos demais professores do Departamento pelos ensinamentos transmitidos;
Aos funcionários técnico-admistrativos, que sempre estão correndo para nos ajudar com
os problemas menores, mas que fazem toda a diferença;
À Grande Rosemar, por toda a simpatia que é, e sempre correndo para agilizar as coisas
para os alunos na seção de pós-graduação. Aos demais funcionários da Seção de pós-
graduação por sempre terem contribuído aos alunos;
Aos professores da graduação, por ter me feito um biólogo e me dar à visão de mundo que
tenho hoje, em especial a Cláudia Carareto e Eleni, por quem tenho grande admiração;
Aos funcionários do Laboratório Síncrotron, pela ajuda nos equipamentos de raios X, em
especial ao Lucas,;
Aos amigos eternos da graduação, que mesmo distante no momento, eu sei que estarão
torcendo por mim, principalmente a Ana Paula, Marcela, Leandra, Aline, Amanda e
Renata e Gustavo;
A todos os demais amigos conquistados durante todos estes anos de IBILCE;
Aos amigos conquistados em Fortaleza, em especial ao Gustavo, Taianá e Vitor, por me
ajudarem quando precisei;
Aos amigos conquistados em São Paulo, pelos ensinamentos transmitidos em “virtual
screen”, em especial, a profa. Dra Antônia Tavares e ao Dr. Humberto;
À Andréia, pela amizade e carinho que tem por mim;
Aos meus amigos de Fernandópolis, que com certeza estarão torcendo por mim, não
importa onde eu ou eles estejam, principalmente ao Ricardo, Rosa, Vagner e Adriana,
Fernando e Miriam, Rogério e Elisângela;
À minha família, em especial a minha querida e amada mãe, do qual sem ela nada sou.
Que é o suporte para toda a minha vida;
À Fapesp pelo auxílio financeiro e concessão da bolsa de doutorado;
A todas as pessoas que de alguma forma contribuíram para a minha formação ou para
a realização deste trabalho, inclusive aquelas que por força do destino não tinham
simpatia por mim ou eu por elas, mas que contribuíram para o meu crescimento pessoal e
profissional;
E finalmente a Deus por minha existência;
“A maior recompensa do nosso trabalho
não é o que nos pagam por ele,
mas aquilo que ele nos transforma”
Joln Ruskin
“-Podes dizer-me, por favor, que caminho devo seguir para sair daqui?
-Isso depende muito de para onde queres ir – respondeu o gato.
-Preocupa-me pouco aonde ir – disse Alice.
-Nesse caso, pouco importa o caminho que sigas – replicou o gato.”
(Lewis Carrol – Alice no país das Maravilhas)
Estudo estrutural de proteínas alvo de Mycobacterium tuberculosis
Lista de Abreviações
μM
micro molar
ACP Acyl carrier protein - Proteína carreadora de acila
ADP Adenosina difosfato
AIDS Adquired immunodeficiency syndrome - Síndrome da Imunodeficiência Adquirida
ATP Adenosina trifosfato
BCG Bacilo Calmette Ghérin
C Carbono
CQ Chiquimato quinase
DAHP Deoxi-D-arabino-heptulosonato-7-fosfato
DHFR Enzima Dihidrofolato redutase
E4P Eritrose-4-fosfato
EPSP 5-enoil-piruvinil-chiquimato-3-fosfato
EPSPS 5-enoil-piruvinil-chiquimato-3-fosfato sintase
ESB Extended shikimate binding -domínio de ligação do chiquimato estendido
FAS I Sistema de síntese de ácidos graxos tipo I
FAS II Sistema de síntese de ácidos graxos tipo II
FMN
Flavina mononucleotídeo
G3P D-gliceraldeído-3-fosfato
HCl Ácido clorídrico
Hepes Ácido N-(2-hidroxietil)piperazina-N’-(2-etanosulfônico)
HIV Human Immunodeficiency virus (Vírus da Imunodeficiêrncia Humana)
IGP Indol glicerol-3-fosfato
INH Isoniazida
InhA Proteína Enoil ACP redutase
IPP
Indol propanol-3-fosfato
KatG Proteína catalase peroxidase
K
i
Constante de inibição
Mg
2+
Íon Magnésio
MtCQ Chiquimato quinase de Mycobacterium tuberculosis
N Nitrogênio
Na Sódio
NAD Nicotinamida adenina difosfato
Estudo estrutural de proteínas alvo de Mycobacterium tuberculosis
NADH-INH Complexo de nicotinamida adenina dinucleotídeo e isoniazida
NADP Nicotinamida adenina dinucleotídeo fosfato
NB Nucleotide binding – referente ao domínio de ligação do nucleotídeo
NMP Mononucleotídeo fosfato
OMS Organização Mundial de Saúde
PDB Protein Data Bank – Banco de dados de proteína
PEG Polietileno glicol
PEP Fosfoenol piruvato
Pi Fosfato inorgânico
PLP Piridoxal fosfato
RC Reduced core – referente ao domínio central reduzido
SB Shikimate binding – referente ao domínio de ligação do chiquimato
SDR Short dehydrogenase/reductase – dehidorgenase/redutase de cadeia curta
SpCS Corismato sintase de Streptococcus pneumoniae
TIM Triose fosfato isomerase
Tris Tris(hidroximetil)aminometano
TRPS Triptofano sintase
RESUMO
A tuberculose representa hoje um dos principais problemas de saúde pública
mundial. É estimado que cerca de cem milhões de pessoas sejam infectadas anualmente.
Para este grande número de casos, dois fatores têm contribuído significativamente, a
resistência da Mycobacterium tuberculosis aos antibióticos existentes e o aumento de
casos de co-infecção com HIV. Por isso, é importante o desenvolvimento de novas
drogas contra o seu agente causador. Há duas abordagens principais para o
desenvolvimento de drogas. Primeiramente pode-se identificar e inibir proteínas que
estejam presentes em uma determinada classe de microorganismos, mas não sejam
conservadas em humanos, levando a antibióticos de largo espectro, ou ainda pode-se
identificar e inibir proteínas de um determinado microorganismo, levando a drogas
específicas. Exemplos da primeira abordagem são as enzimas da via do ácido
chiquímico e suas ramificações, como a via de síntese de triptofano. A via do ácido
chiquímico é responsável pela biossíntese de corismato, que é o precursor comum
utilizado na geração de vários compostos aromáticos importantes para sobrevivência da
bactéria, entre eles os aminoácidos e co-enzimas; e exemplos da segunda abordagem
são enzimas relacionadas com a síntese de ácidos micólicos, que são importantes
constituintes da parede celular de micobactérias. O objetivo do presente trabalho foi
realizar estudos estruturais de quatro proteínas que são alvos em potencial para o
desenvolvimento de drogas contra tuberculose, sendo elas: a Chiquimato quinase e
Corisamto sintase, ambas da via do ácido chiquímico; Triptofano sintase, última enzima
da via de síntese de triptofano; e a enzima dependente de NADH enoil ACP redutase
(InhA), relacionada com a síntese de ácidos micólicos. Neste trabalho são apresentadas
as estruturas da Chiquimato quinase de M. tuberculosis em complexo com ADP e
magnésio, na ausência do substrato ácido chiquímico, e da Chiquimato quinase em
complexo com ADP e ácido chiquímico, na ausência do íon magnésio. Nestas estruturas
pôde-se observar o efeito de íons e a influência do ácido chiquímico sobre a estrutura da
chiquimato quinase. Apresentamos também, a primeira estrutura da corismato sintase de
Mycobacterium tuberculosis, evidenciando suas características estruturais. Além disso,
foram realizados estudos estruturais com a proteína InhA na sua forma nativa,
mostrando o seu estado conformacional na ausência de ligantes, e três estruturas para
esta proteína em complexo com isoniazida ativa, sendo a proteína selvagem e dois
mutantes identificados clinicamente como resistentes a este medicamento (S94A e
I21V). Além desses estudos, foi realizada a modelagem molecular da proteína triptofano
sintase em complexo com seis inibidores, e feita análise da especificidade destes com
relação a esta proteína. Com a realização deste trabalho pretendemos estar contribuindo
para a solução do grande problema que a tuberculose representa hoje, por afligir
milhões de pessoas em todo o mundo.
ABSTRACT
Tuberculosis represents today a large problem of world public health. It is
estimated that approximately a hundred million people are infected annually. For this
large number of cases, two factors have contributed significantly, the resistance of
Mycobacterium tuberculosis to existent antibiotics and the increase of the cases of the
co-infection with HIV. Therefore, the development of new drugs against its etiologic
agent is important. There are two main approaches for the development of drugs.
Firstly, proteins of a determined class of microorganism, that are not present in humans,
can be identified and inhibited leading to large-spectra antibiotics; or it can identify and
inhibit proteins of a determined organism leading to specific drugs. Examples of the
first approach are enzymes of the shikimate pathway and its branches, such as the
tryptophan synthesis pathway. The shikimate pathway is responsible for biosynthesis of
chorismate, it is a common precursor utilized in the production of various important
aromatic compounds used in the survival of the bacteria, such as amino acids and co-
enzymes; an example of the second approach are enzymes related to the synthesis of
mycolic acids, they are important constituents of the cellular wall of mycobacteria. The
objective of the present work was to perform structural studies of four enzymes that are
potential targets to the development of drugs against tuberculosis, such as: shikimate
kinase and chorismate synthase, both of the shikimate pathway; tryptophan synthase,
last enzyme of the trytophan pathway; and the dependent of NADH enzyme enoyl ACP
reductase (InhA) related to the synthesis of mycolic acids. Here, the shikimate kinase
structures from M. tuberculosis in complex with ADP and magnesium ion, in the
absence of shikimate and the shikimate kinase in complex with ADP and shikimate, in
the absence of the magnesium ion, are presented. The effects of ions and shikimate on
the structure of shikimate kinase were observed. Here, we also present the first structure
of chorismate synthase from M. tuberculosis showing its structural characteristics.
Furthermore, we perform structural studies of the InhA protein in its native form
showing its conformational state in the absence of ligands and three structures for this
protein in complex with active isoniazid (wild type and two resistant clinical isolated
mutants to this medicine, S94A and I47T). We also performed the molecular modeling
of the tryptophan synthase protein in complex with six inhibitors, and the analysis of the
specificity of these in relation to the protein was performed. With the completion of this
work, we intend to contribute to the solution of the big program that tuberculosis
represents today causing suffering to the millions of people worldwide.
Estudo Estrutural de Enzimas Alvo de Mycobacterium tuberculosis
ÍNDICE DE FIGURAS
Figura 1 Mycobacterium tuberculosis corado pelo método ácido-resistente ..................... 3
Figura 2 Representação esquemática da via metabólica do ácido chiquímico ................. 5
Figura 3 Representação esquemática das ramificações da via do ácido chiquímico ....... 7
Figura 4 Reação catalisada pela chiquimato quinase ......................................................... 9
Figura 5 Representação esquemática da chiquimato quinase ........................................... 10
Figura 6 Domínios estruturais para a chiquimato quinase ................................................ 11
Figura 7 Alterações estruturais e efeito sinérgico observado na chiquimato quinase ..... 13
Figura 8 Reação catalisada pela corismato sintase ............................................................. 15
Figura 9 Representação esquemática da corismato sintase ........................ ...................... 15
Figura 10 Completo mecanismo catalítico para a corismato sintase .................................. 17
Figura 11 Representação esquemática da via de síntese de triptofano ............................... 20
Figura 12 Reação catalisada pela triptofano sintase ............................................................. 21
Figura 13 Estrutura tridimensional da triptofano sintase ................................................... 22
Figura 14 Formula molecular dos ligantes para triptofano sintase .................................... 24
Figura 15 Representação esquemática do envelope celular de micobactérias ................... 25
Figura 16 Reação catalisada pela enzima InhA .................................................................... 26
Figura 17 Representação esquemática da InhA .................................................................... 27
Figura 18 Mecanismo de ativação da isoniazida .................................................................. 28
Figura 19 Contatos moleculares entre o isonicotínico acil-NADH e o sítio ativo da InhA 29
Estudo estrutural de proteínas alvo de Mycobacterium tuberculosis
SUMÁRIO
1. INTRODUÇÃO
1
1.1 A tuberculose e a necessidade do desenvolvimento de novas drogas
1
1.2 A via do ácido chiquímico
5
1.3 A chiquimato quinase
8
1.4 A corismato sintase
14
1.5 A triptofano sintase
18
1.6 InhA e o mecanismo de resistência ao antibiótico isoniazida
24
2. OBJETIVOS
30
3. RESULTADOS
31
3.1 Crystallization and preliminary X-ray crystallographic analysis of
chorismate synthase from Mycobacterium tuberculosis
32
3.2 Structure of chorismate synthase from Mycobacterium tuberculosis 36
3.3 Molecular models of tryptophan synthase from Mycobacterium tuberculosis
complexed with inhibitors
52
3.4 Effects of the magnesium and chloride ions and shikimate on the structure
of shikimate kinase from Mycobacterium tuberculosis
64
Estudo estrutural de proteínas alvo de Mycobacterium tuberculosis
3.5 Crystallographic studies on the binding of isonicotinyl-NAD adduct to wild-
type and isoniazid resistant 2-trans- Enoyl-ACP (CoA) Reductase from
Mycobacterium tuberculosis
72
4. CONCLUSÕES
118
5. BIBLIOGRAFIA
121
6. ANEXO – Artigo de revisão publicado
126
6.1 Chorismate Synthase: an attractive target for drug development against
neglected disease
127
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
1. INTRODUÇÃO
1.1 A tuberculose e a necessidade do desenvolvimento de novas drogas
As doenças infecciosas é a quarta maior causa de morte da população humana
(KING et al., 2006), sendo responsável pelo sofrimento e morte de centenas de milhões de
pessoas, especialmente em áreas tropicais e subtropicais do mundo, local onde ocorre cerca
de 90% das mortes causadas por esse tipo de enfermidade (TROUILLER et al., 2001).
Segundo a Organização Mundial de Saúde (OMS) cerca de 14 milhões de pessoas morrem
a cada ano devido a desse tipo de doença, destacando-se entre elas AIDS/HIV, infecções
respiratórias, malária e tuberculose, que lideram a causa de morte e de morbidade,
principalmente nos países da África, Ásia e América do Sul – regiões nas quais estão
presentes oitenta por cento de toda a população mundial (TROUILLER et al., 2001).
A organização mundial de saúde identifica três fatores chaves que podem contribuir
coletivamente para a pandemia associada às doenças infecciosas. São eles: falência no uso
de ferramentas existentes no controle de tais doenças, ferramentas inadequadas ou
inexistentes, e conhecimento insuficiente sobre a doença (TROUILLER et al., 2002).
Dentre essas doenças, a tuberculose, hoje, é considerada a causa infecciosa líder de
mortes. É estimado que cem milhões de pessoas são infectadas anualmente. De cem
milhões, aproximadamente dez milhões desenvolvem a doença e três milhões morrem
(MORGAN; HARITAKUL; KELLER, 2003).
A incidência de tuberculose teve um declínio rápido no início do século vinte nos
países em desenvolvimento, devido à melhora nas condições sanitárias e de moradias. Essa
tendência foi, inicialmente, acelerada pela introdução da vacinação BCG (1927) e da
descoberta de antibióticos como a estreptomicina (1944) e, posteriormente, com a
descoberta do ácido p-aminosalisílico (1946), isoniazida (1952) e rifampicina (1965)
(DUNCAN, 2003). Mas está ocorrendo uma ressurgência dos casos de tuberculose nos
1
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
países em desenvolvimento. Muitos são os fatores envolvidos, porém, o mais notável é a
falta de recursos e o descaso governamental que impedem a implementação adequada de
medidas de controle. Além disso, a incidência aumentou também nos países desenvolvidos,
principalmente devido à imigração e à negligência direta do mundo todo com relação ao
problema (RAVIGLIONE et al., 1997). Todavia, dois recentes eventos têm contribuído
significativamente para a pandemia de tuberculose: a) o aumento de bactérias multi-
resistentes devido a terapias inadequadas e o uso indiscriminado de antibióticos,
principalmente a isoniazida e rifampicina, as duas drogas mais utilizadas no tratamento
quimioterápico (BAPTISTA et al., 2002; DUNCAN, 2003); e b) a AIDS, pois pessoas
portadoras de ambos os agentes causadores da tuberculose e AIDS têm o sistema imune
debilitado, e, portanto, maiores chances de desenvolver a doença (BAPTISTA et al., 2002;
DYE et al., 2002). Em 1993, a Organização Mundial de Saúde declarou a tuberculose
como uma emergência global. Entretanto, a incidência exata desta doença no panorama
mundial não é bem conhecida (RAVIGLIONE et al., 1997).
O principal responsável pela tuberculose em humanos é a bactéria Mycobacterium
tuberculosis. Este organismo é um bastonete fino, com comprimento médio de 4 μm, como
pode ser visto pela coloração por métodos ácido-resistentes (Fiehl-Neelsen) (Figura 1) ou
fluorescentes (COTRAN et al., 1991). Apresenta crescimento lento, dormência e complexo
envelope celular. É um organismo obrigatoriamente aeróbico e é um parasita intracelular
facultativo, usualmente de macrófagos. O tempo de geração da M. tuberculosis em animais
infectados é de aproximadamente vinte e quatro horas. Isto contribui para a característica
crônica da doença (apesar da doença em si não apresentar uma fase crônica),
impossibilitando fortemente o regime de tratamento e representando um formidável
obstáculo para pesquisas (COLE et al., 1998). M. tuberculosis é quase exclusivamente um
parasita humano e é transmitido pelo ar. Usualmente causa tuberculose pulmonar, uma
2
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
infecção dos pulmões, embora, seja capaz de infectar outros tecidos ou órgãos do corpo
(PELCZAR et al., 1996).
Figura 1. Mycobacterium tuberculosis corado
pelo método ácido-resistente (Fiehl-Neelsen).
Clinicamente a tuberculose pode ser classificada como primária e secundária.
Sendo que a tuberculose primária é caracterizada por uma forma de doença que se
desenvolve previamente numa pessoa não-exposta e, portanto, não-sensibilizada. A
tuberculose primária inicia-se pela inalação do M. tuberculosis e culmina com o
desenvolvimento da imunidade mediada por células no organismo. Na maioria das pessoas,
a tuberculose primária é assintomática, ainda que possa causar febre e derrame pleural. Das
pessoas recém-infectadas somente 5% desenvolvem a doença em uma forma clinicamente
significante. A tuberculose secundária é um padrão de doença que observa em um
hospedeiro previamente sensibilizado. Pode surgir pouco depois da tuberculose primária,
porém mais frequentemente, ela surge da reativação das lesões primárias latentes muitas
décadas após a infecção inicial, particularmente quando a resistência do hospedeiro estiver
enfraquecida. No caso da tuberculose pulmonar secundária, ela é geralmente localizada no
ápice dos lóbulos superiores de um ou de ambos os pulmões. A tuberculose secundária
localizada pode ser assintomática. Quando as manifestações aparecem, elas são,
geralmente, insidiosas no início. Os sintomas sistêmicos aparecem, com freqüência no
início do curso e incluem indisposição, anorexia, perda de peso e febre. Com o
3
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
envolvimento pulmonar progressivo, surgem quantidades cada vezes maiores de escarro,
primeiramente mucóide e posteriormente purulento. As manifestações extra pulmonares da
tuberculose são em grande número e dependem do sistema orgânico envolvido (KUMAR,
ABBAS, FASTO, 1999)
Com a determinação completa do genoma da M. tuberculosis ocorreu um profundo
avanço em pesquisas sobre tuberculose (COLE et al., 1998). Novas abordagens e técnicas
têm sido utilizadas para avaliar as informações obtidas pelo projeto genoma deste
organismo. Entre elas, inclui a utilização de ferramentas de bioinformática, proteômica,
transcriptoma e genômica estrutural combinadas com ferramentas de manipulação de
genoma. Com a utilização destas técnicas tem-se permitido um melhor entendimento da
complexa biologia deste patógeno (DUNCAN, 2003). Além disso, a determinação do
genoma e o melhor esclarecimento dos mecanismos de virulência deste organismo,
permitem a possibilidade da descoberta de novos alvos para o desenvolvimento de drogas
(COLE et al., 1998; DUNCAN, 2003; KHASNOBIS; ESCUYER; CHATTERJEE, 2002).
Para a identificação de alvos com potencial de gerar novas drogas contra M.
tuberculosis pode se utilizar duas abordagens significativas. A primeira, em que se podem
isolar proteínas que são conservadas entre diferentes espécies de bactérias, mas que não
conservadas em humanos, conduzindo aos antibióticos de largo espectro; e a segunda na
qual se podem identificar alvos que são únicos em M. tuberculosis, levando a obtenção de
drogas anti-tuberculose específicas (KHASNOBIS; ESCUYER; CHATTERJEE, 2002).
Exemplos da primeira abordagem são as enzimas que compõe a via de síntese de
ácido corísmico (via do ácido chiquímico) e da segunda abordagem são as enzimas
responsáveis pela síntese de ácidos micólicos.
4
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
1.2 A Via do Ácido Chiquímico
A via do ácido chiquímico (Figura 2) é essencial para algas, plantas, bactérias,
fungos (BENTLEY, 1990) e parasitas do filo apicomplexa, como Plasmodium,
Toxoplasma e Cryptosporidium (ROBERTS et al., 1998), e não está presente em
mamíferos.
Figura 2. Representação esquemática da via metabólica do ácido chiquímico.
5
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
A via do ácido chiquímico é definida como os sete passos metabólicos iniciais da
condensação de fosfoenol piruvato (PEP) e eritrose-4-fosfato (E4P) finalizando com a
síntese de corismato. Os sete passos desta via foram originalmente descobertos através de
estudos em bactéria, principalmente Escherichia coli e Salmonella typhimurium
(HERRMANN; WEARVER, 1999).
A via do ácido chiquímico foi inicialmente descrita como a rota biossintética para a
produção de aminoácidos aromáticos (Fenilalanina, Tirosina e Triptofano) através dos
estudos clássicos de Bernhard Daves e David Sprinson e seus colaboradores por volta de
1950. A identificação de seus vários intermediários foi completada no início de 1960,
sendo que o primeiro deles a ser identificado foi o ácido chiquímico (PITTARD, 1987;
HERMANN, 1995). Em microorganismos, esta via, além de ser utilizada para síntese de
aminoácidos aromáticos, também faz parte da síntese de folato, quinonas e alguns
metabólitos secundários, como sideróforos e micobactinas. A via do ácido chiquímico
ramifica-se em muitos pontos, porém, corismato é o último ramo comum da via para os
compostos citados acima. Corismato é convertido por cinco enzimas distintas para
prefenato, antranilato, aminodeoxicorismato, isocorismato e p-hidroxibenzoato. Esses
metabólitos compreendem os primeiros intermediários na biossíntese de fenilalanina e
tirosina, triptofano, folato, menaquinonas e sideróforos, e ubiquinonas, respectivamente
(Figura 3) (DOSSELAERE; VANDERLEYDEN, 2001). Sendo que destes intermediários,
folato é um composto que participa de várias formas na síntese de purinas, timidilato,
metionina, glicina e pantotenato; Sideróforos são moléculas responsáveis pela captação de
ferro por bactérias; Quinonas são moléculas lipofílicas, componentes não protéicos da
cadeia transportadora de elétrons ligados à membrana que podem ser divididas em dois
grandes grupos estruturais: benzoquinonas e naftoquinonas. As benzoquinonas são
denominadas de ubiquinonas e as naftoquinonas são chamadas de menaquinonas
(DOSSELAERRE; VANDERLEYDEN, 2001).
6
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figura 3. Representação esquemática das ramificações da via do ácido chiquímico.
(1) corismato piruvato-liase; (2) isocorismato sintase; (3) aminodeoxicorismato
sintase; (4) antranilato sintase; (5) corismato mutase.
A análise completa da seqüência do genoma da M. tuberculosis revelou a presença
dos sete genes aro envolvidos na biossíntese de corismato. São eles, aro D, B, K, F, G, E e
A. Estes genes são responsáveis pela síntese das enzimas 3-dehidroquinato dehidratase, 3-
dehidroquinato sintase, chiquimato quinase, corismato sintase, DAHP sintase, chiquimato-
5-dehidrogenase e EPSP sintase, respectivamente (Figura 2) (COLE et al., 1998; PARISH;
STOKER, 2002).
7
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
A importância desta via, tanto para M. tuberculosis como para outros patógenos foi
justificada por diversos trabalhos: Parish e Stoker (2002) demonstraram que a via do ácido
chiquímico é essencial para M. tuberculosis, pois, eles foram incapazes de isolar cepas
contendo o gene aroK (responsável pela produção da chiquimato quinase) inativado. A
inativação do gene aroD (responsável pela produção de 3-dehidroquinato dehidratase)
também é usada com sucesso para gerar cepas atenuadas para a vacina oral contra
Salmonella typhi e outras bactérias (TACKET et al., 1997). Os estudos de inibição
utilizando glifosato demonstram que este herbicida inibe a EPSP sintase, e apresenta uma
significativa inibição no crescimento de T. gondii, P. falciparum e C. parvum (ROBERTS
et al., 2002).
1.3 A Chiquimato quinase
A Chiquimato quinase, principalmente de Mycobacterium tuberculosis (peso
molecular de 18583,3 Da), é uma enzima que teve sua estrutura intensivamente estudada a
partir de 2002 (GU et al., 2002, DHALIWAL et al., 2004, PEREIRA et al., 2004, GAN et
al., 2006, HARTMANN et al., 2006). Esta proteína, que é, a quinta enzima da via do
ácido chiquímico, catalisa a fosforilação do grupo 3-hidroxil do ácido chiquímico, usando
ATP como um co-substrato (Figura 4) (HERRMANN, 1995) e os íons magnésio e cloreto
como co-fatores. Em Escherichia coli, esta reação é catalisada por duas diferentes
isoformas, sendo elas a CQ I, codificada pelo gene aroK e a CQ II codificada pelo gene
aroL. Estas proteínas são monoméricas, possuem uma massa molecular de
aproximadamente 19.500 Dáltons e apresentam uma identidade de 30% entre si. Ambas as
enzimas possuem atividade in vitro na biossíntese de aminoácidos aromáticos. Mas in
vivo, a isoenzima II não tem uma atividade exclusiva da via do ácido chiquímico, pois
parece apresentar também uma função na divisão celular (HERRMANN; WEAVER,
1999; 1995; KRELL; COGGINS; LAPTHORN, 1998). Diferente de E. coli, o genoma
8
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
completo de algumas bactérias, como Mycobacterium tuberculosis e Haemophilus
influenzae, tem revelado a presença de apenas um gene para chiquimato quinase. Muitas
destas chiquimato quinases parecem ser codificadas pelo gene aroK, pois sua seqüência de
aminoácidos tem alta identidade com a CQ I de E. coli (GU et al., 2002).
Chiquimato 3-fosfato
Ácido chiquímico
Figura 4. Reação catalisada pela chiquimato quinase.
A primeira estrutura cristalográfica da chiquimato quinase de M. tuberculosis
(MtCQ) foi resolvida na presença de Mg-ADP (GU et al., 2002). A MtCQ é uma proteína
da classe α/β que apresenta na sua estrutura fitas β rodeadas por hélices α. No centro da
proteína estão localizadas cinco fitas β paralelas (β1-β5) formando uma folha β na ordem
23145, que é flanqueada por oito hélices (Figura 5) (GU et al., 2002).
9
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
α
1
β5
β4
β3
β2
β1
Figura 5. Represenção de elementos de estrutura secundária da MtCQ em complexo com Magnésio
(Mg
2+
), Adenosina difosfato (ADP) e ácido chiquímico (PEREIRA et al., 2004) gerada pelo programa
MolMol (KURADI; BILLETER; WÜTHRICH, 1996).
A ordem da folha β 23145 observada na estrutura da MtCQ classifica esta proteína
como pertencente à família das monofosfato nucleosídeo (NMP) quinases (YAN; YSAI,
1999). As proteínas desta família são compostas por três domínios: central (local de
ligação do fosfato), domínio da tampa (Lid domain) e o domínio de ligação do NMP ou SB
(shikimate binding). A MtCQ possui uma seqüência de aminoácidos conservada,
GXXXXGKT/S, conhecida como motivo Walker A, que forma o P-loop (SCHULZ,
1992). Esta região liga-se ao β-fosfato presente no ADP e está localizada entre a β1 e α1.
As NMP quinases sofrem grande mudança conformacional durante a sua atividade
catalítica. As regiões responsáveis por este movimento estão localizadas no sítio de ligação
do NMP e no Lid domain (MÜLLER et al., 1996; GERSTEIN; SCHULZ; CHOTHIA,
1993).
Hartmann et al. (2006) baseado na análise de movimentos globais acompanhados
no processo de ligação de substratos no sítio ativo da chiquimato quinase, reclassificou os
10
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
domínios da chiquimato quinase. Segundo sua classificação, temos agora quatro domínios,
que são: o ESB (extended SB) que inclui o subdomínio SB; o NB (nucleotide-binding) que
inclui o P-loop e o AB-loop ( adenine binding loop) e um curto seguimento da hélice α 6; o
LID; e a parte remanescente que forma o RC (reduced core) (Figura 6). O antigo domínio
central compreende os presentes domínios RC e NB e um curto seguimento do domínio
ESB.
Figura 6. Domínios estruturais para a Chiquimato quinase proposto por
Hartmann et al. (2006). NB: nucleotide-binding, RC: reduced core, ESB:
extended Shikimate binding, LID: Lid domain
Apesar dos estudos cristalográficos feitos com as chiquimato quinases de M.
tuberculosis (GU et al., 2002), E. coli (ROMANOWSKI; BURLEY, 2002) e Erwinia
chrysanthemi (KRELL; COGGINS; LAPTHORN, 1998) a posição precisa do ácido
chiquímico só foi determina em 2004 por Pereira et al. (2004) e Dhaliwal et al. (2004). A
determinação da estrutura da chiquimato quinase em complexo com ácido chiquímico
11
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
revelou os resíduos do sítio ativo que participam das interações entre este substrato e a
enzima. Pela análise da estrutura deste complexo puderam-se observar também as
mudanças conformacionais ocasionadas pela presença do substrato em seu sítio ativo. As
principais alterações conformacionais ocorrem no domínio ESB e no Lid domain, na qual
ambos movem-se em direção ao centro da enzima causando um fechamento parcial do sítio
ativo da enzima (PEREIRA et al., 2004; DHALIWAL et al., 2004).
Entretanto, muito recentemente Gan et al. (2006) e Hartmann et al. (2006)
mostraram através de dados estruturais que a molécula de nucleotídeo é o principal
responsável pelo fechamento do Lid domain sobre o sítio ativo da MtCQ. Hartmann et al.
(2006) mostrou que ocorre um sinergismo de ligação dos substratos no sítio ativo desta
enzima. Segundo estes autores, ocorrem um efeito sinérgico aleatório de ligação tanto para
o ácido chiquímico como para o nucleotídeo. Esse efeito sinérgico pode ser consistente
com o possível aumento da afinidade para ambos os substratos, independente da ordem de
ligação. A Figura 7 mostra o efeito sinérgico de ambos os substratos e suas influências na
estrutura da chiquimato quinase de Mycobacterium tuberculosis proposto por Hartmann et
al. (2006).
12
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figura 7. Alterações conformacionais observadas na Chiquimato quinase de M.
tuberculosis na presença de seus ligantes e o possível efeito sinérgico desses ligantes
segundo Hartmann et al. (2006). SKM: ácido chiquímico; ATP: adenosina
trifosfato
Até muito recentemente, não se conhecia a conformação da estrutura da chiquimato
quinase na ausência de qualquer ligante. Entretanto Gan et al. (2006) e Hartmann et al.
(2006) determinaram a estrutura desta enzima no seu estado nativo que possibilita intuir os
movimentos causados pela molécula de ADP. Esses autores também determinaram a
estrutura da chiquimato quinase em complexo com análogo de ATP (GAN et al., 2006) e
ATP (HARTMANN et al., 2006) e, desta forma, pode-se acrescentar informações sobre o
mecanismo de transferência do fosfato da molécula de ATP para a molécula de ácido
chiquímico.
13
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Com relação ao íon cloreto que interage com a chiquimato quinase, foi observada a
influência deste sobre a proteína apenas por técnicas de baixa resolução (CERASOLI et al.,
2003). Assim, por meio de dicroísmo circular e de cinética enzimática foi observado que a
presença do íon cloreto enfraquece a interação entre o ácido chiquímico e a chiquimato
quinase de Erwinia Chrysanthemi, mas, por outro lado, fortalece a afinidade da enzima por
ADP ou ATP (CERASOLI et al., 2003). Entretanto, apesar deste estudo, pouco se sabe
sobre o papel deste íon e do íon magnésio sobre a estrutura tridimensional desta proteína.
1.4 A Corismato Sintase
Outra enzima da via do ácido chiquímico, que a partir de 2004 se tornou alvo de
muitas pesquisas na área de biologia estrutural, é a corismato sintase (AHN et al., 2004,
VIOLA; SARIDAKIS; CHRISTENDAT et al., 2004, MACLEAN; ALI et al., 2003,
QUEVILLON-CHERUEL et al., 2004, DIAS et al., 2006). Esta proteína é a última enzima
da via do ácido chiquímico e realiza a conversão de 5-enoilpiruvil-3-chiquimato fosfato
(EPSP) em corismato, através de uma reação de eliminação de um grupo fosfato da
molécula de EPSP (Figura 8) (KITZING; MACHEROUX; AMRHEIN, 2001). A
eliminação desse grupo fosfato é responsável pela adição da segunda dupla ligação na
molécula de corismato (PITTARD, 1987). A corismato sintase requer para sua reação a
presença de flavina mononucleotídeo (FMN), embora a reação não envolva um estado de
mudança redox (HERRMANN, 1995). A eliminação do grupo fosfato é realizada por um
mecanismo não usual, através de uma 1,4-anti-eliminação. Até agora não foi descrita
nenhuma outra enzima que catalisa uma reação deste tipo. Sabe-se também, que na reação
está envolvido um intermediário instável de vida curta, formado depois da perda do fosfato
da molécula de EPSP. Estudos recentes têm dado evidência de que esse intermediário é um
radical, formado pela transferência de um elétron da molécula de FMN para o EPSP
(MACLEAN; ALI, 2004; KITZING, 2004).
14
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figura 8. Reação catalisada pela corismato sintase.
A reação catalisada pela corismato sintase ocorre em duas etapas, em que,
primeiramente ocorre a clivagem da ligação C3-O3 e o lançamento do fosfato, produzindo
o intermediário instável, que colapsa para formar o corismato no segundo passo devido à
clivagem da ligação 6proR-C-H. A orientação da molécula de FMN na estrutura da
corismato sintase sugere que ela pode promover a estabilização de qualquer intermediário
deficiente em elétron da molécula do EPSP (Figura 9) (MACLEAN; ALI, 2004).
Figura 9. Representação de elementos de estrutura secundária da corismato sintase de
Streptococcus pneunoniae complexada com 5-enoilpiruvil-chiquimato 3-fosfato (EPSP)
e flavina mononucleotídeo (FMN) (MACLEAN; ALI 2004), gerada pelo programa
MolMol (KURADI; BILLETER; WÜTHRICH, 1996).
15
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Corismato sintases de diferentes organismos podem ser classificadas de acordo com
sua capacidade em reduzir a molécula de flavina mononucleotídeo. Desta forma, temos as
corismato sintases monofuncionais e as bifuncionais. Corismato sintases de fungos
apresentam uma segunda função, além da conversão de EPSP em corismato. Elas são
também capazes de realizar a redução da molécula de FMN utilizando NADH. Assim,
estas corismato sintases são classificadas como bifuncionais, por apresentarem também
uma função intrínseca de flavina redutase (MACHEROUX et al., 1999). Por outro lado, as
corismato sintases de plantas, bactérias e parasitas do filo apicomplexa não possuem essa
capacidade de redução de flavina mononucleotídeo, tendo que obter esta molécula na sua
forma reduzida do meio e, portanto, são classificadas como monofuncionais
(MACHEROUX et al., 1999).
Apesar da estrutura cristalográfica da corismato sintase de vários organismos terem
sido resolvidas (AHN et al., 2004, VIOLA; SARIDAKIS; CHRISTENDAT et al., 2004,
MACLEAN; ALI et al., 2003, QUEVILLON-CHERUEL et al., 2004), somente a estrutura
de Streptococcus pneumoniae (SpCS) apresenta ambas moléculas de flavina
mononucleotídeo (FMN) e o substrato 5-enoilpiruvil-3-chiquimato fosfato (EPSP) ligadas
no seu sítio ativo (MACLEAN, ALI, 2003). Desta forma, esta estrutura revelou muitas
informações sobre o complexo mecanismo catalítico da corismato sintase. Baseado nesta
estrutura e na análise de proteínas mutantes de Neurospora Crasa, nas quais dois
importantes resíduos de histidina do sítio ativo foram mutados para alanina, pôde-se propor
um mecanismo mais completo para a reação catalisada por esta enzima. Segundo este
mecanismo, após a ligação da molécula de EPSP no sítio ativo da enzima com a molécula
de FMN reduzida já ligada, um elétron é transferido para a dupla ligação do substrato,
iniciando desta maneira, a clivagem e o lançamento do fosfato, com um resíduo de
histidina atuando como um neutralizador da carga incipiente do átomo de oxigênio (Figura
10). O resultante carbono neutro C(4a) da flavina semiquinona tautomeriza-se para uma
16
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
espécie radial na qual o elétron não pareado reside sobre o átomo N(5), que
concomitantemente causa a abstração do hidrogênio ligado ao C(6). E finalmente, ocorre a
deprotonação da flavina reduzida restaurando o seu estado inicial (Figura 10) (KITZING et
al., 2004).
Figura 10. Mecanismo catalítico completo para a Corismato Sintase proposto
por Kitzing et al. (2004). Rib: Ribose
17
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Apesar das estruturas cristalográficas da corismato sintase de Aquiflex aeolicus
(VIOLA; SARIDAKIS; CHRISTENDAT et al., 2004) e S. cerevisiae (QUEVILLON-
CHERUEL et al., 2004) estarem na forma nativa, grande parte de suas estruturas não foi
resolvida, devido à ausência de densidade eletrônica, principalmente alguns importantes
resíduos do sítio ativo. Assim, a conformação desses no sítio ativo da corismato sintase na
ausência de qualquer ligante é desconhecida.
A estrutura da corismato sintase apresenta-se como um tetrâmero, onde cada
monômero possui um único domínio central, que é rodeado por loops e trechos discretos
de hélices α e folhas β. O centro do monômero consiste de uma camada de quatro longas
hélices α, sanduíche entre um par de 4 fitas de folhas β antiparalelas (MACLEAN; ALI,
2003) (Figura 9). O bolsão de ligação apresenta uma parte hidrofóbica, na qual está
localizado o sistema de anéis isoaloxazino da molécula de FMN e uma outra parte
hidrofílica carregada positivamente devido a grande quantidade de argininas e histidinas.
Nesta segunda parte do sítio, está localizado a porção remanescente da molécula de FMN e
o EPSP, na qual estas duas moléculas apresentam-se justapostas a uma distância
aproximada de 3,0 Å, favorecendo, assim, a transferência do elétron entre a molécula de
EPSP e FMN (MACLEAN; ALI, 2004).
1.5 A Triptofano Sintase
As ramificações da via do ácido chiquímico também representam alvos
interessantes para o desenvolvimento de drogas de amplo espectro e vacinas para doenças
causadas por microrganismos, pelo fato de também estarem ausentes em mamíferos.
Exemplos clássicos, é a inibição da dihidrofolato redutase (DHFR), terceira enzima da via
de síntese do ácido p-aminobenzóico, por trimetoprim (HAWSER; LOCIURO; ISLAM,
2006) e a inibição da dihidropteroato sintase, quarta enzima da mesma via, pelos
antibióticos da série das sulfonamidas (SKÖLD, 2001). Outros exemplos de alvo são as
18
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
enzimas da via de síntese de triptofano, como demonstrado por Smith (2001), em que a
inativação do gene trpD, responsável pela síntese da proteína antranilato
fosforibosiltransferase, produz cepas atenuadas de M. tuberculosis. A inativação do gene
trpA (corresponde a subunidade I da antranilato sintase) também foi demonstrado gerar
cepas atenuadas de Bordetella bronchiseptica e Leptospira meyeri (McARTHUR et al.,
2003; BAUBY; GIRONS; PICARDEAU, 2003). A enzima triptofano sintase demonstrou
ser alvo para o desenvolvimento de herbicidas, pois testes in vitro com inibidores análogos
de substrato demonstraram ter atividade inibitória para esta enzima (FINN et al., 1999).
Desta forma a via de síntese de triptofano é bastante estudada no desenvolvimento
de terapias contra doenças infecciosas. Esta via é composta por cinco enzimas, sendo elas,
a antranilato sintase, antranilato fosforibosiltransferase, fosforibosil-antranilato isomerase,
indol-glicerol-3-fosfato sintase e a triptofano sintase (Figura 11).
19
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figura 11. Via de síntese do aminoácido triptofano
Dentre estas enzimas, uma das mais bem estudadas é a triptofano sintase, sendo
muito bem conhecido os detalhes de sua purificação, bioquímica, inibição e modificações
estruturais de sua atividade (PAN; WOEHL; DUNN, 1997).
A triptofano sintase é responsável pela catálise das duas últimas reações da
biossíntese de L-triptofano. Esta proteína é uma enzima bifuncional com duas subunidades,
denominadas α e β. A subunidade α catalisa uma reação retroaldol em que o indol-3-
20
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
glicerol fosfato (IGP) é clivado em indol e D-gliceraldeído-3-fosfato (G3P). O indol
produzido pela subunidade α reage com L-serina ativada por uma reação dependente de
piridoxal fosfato (PLP) no sítio ativo da subunidade β para formar L-triptofano e água
(Figura 12A, B).
Indol
B
A
Figura 12. Reação catalisada pela triptofano sintase. A) Reação catalisada pela
subunidade α. B) Reação catalisada pela subunidade β. IGP: Indol glicerol 3-fosfato;
G3P: Glicerol fosfato; L-Ser: L-Serina; Enz: Enzima; E(Ain): Almidina interna; E(A-
A
)
: Aminoacrilato
;
E
Q
: Es
p
écie
q
uinóide
;
L-tr
p
: L-tri
p
tofano.
21
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
A subunidade α da triptofano sintase apresenta um enovelamento similar a da
proteína triose fosfato isomerase (TIM) (BANNER et al., 1975), apresentando uma
estrutura α/β barril (MILES, 1995). Entretanto, a subunidade β apresenta dois domínios de
aproximadamente mesmo tamanho, sendo eles denominados N-domínio e C-domínio
(HYDE et al, 1988). O sítio ativo da subunidade α é localizado próximo à interface com a
subunidade β em uma depressão que permite interações de hidrogênio com o grupo fosfato
do substrato IGP e um ambiente hidrofóbico para a sua porção indol (WEYAND, M.;
SCHILICHTING, 1999). O sítio ativo da subunidade β apresenta-se enterrado no centro
desta subunidade entre os N- e C- domínios. A molécula de PLP ligada ao sítio ativo desta
subunidade forma uma ligação covalente tipo C=N (base de Schiff) com um grupo amino
de um resíduo de lisina (Figura 12) (HYDE et al, 1988).
Figura 13. Estrutura tridimensional da Triptofano sintase de Salmonella typhimurium
em complexo com Indol glicerol 3-fosfato (IGP) na subunidade α e piridoxal fosfato
(PLP) na subunidade β (HYDE et. al, 1988).
22
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
A estrutura tridimensional da triptofano sintase de Salmonella typhimurium foi a
primeira a ser resolvida e revelou a presença de um longo canal hidrofóbico que conecta os
sítios ativos da subunidade α e da subunidade β. Este canal teria o papel de impedir a
difusão do substrato e do produto da reação catalisada pela proteína, e também de estar
envolvido na regulação alostérica que sincroniza as reações nas duas subunidades
(SACHPATZIDIS et al., 1999). Após a determinação da estrutura nativa e de outras
estruturas com diferentes ligantes para S. typhimurium, pode-se observar que ligantes
induzem mudanças conformacionais que são importantes para o posicionamento do indol
dentro do canal e também para a comunicação alostérica entre as duas subunidades.
Drásticas mudanças conformacionais ocorrem com a adição de ligantes, tanto na
subunidade α como na subunidade β. Estas mudanças incluem ordenação de loops
próximos ao sítio ativo da subunidade α e movimentos de subdomínios da subunidade β
(MILES, RHEE, DAVIES, 1999).
Entre os ligantes estudados que se ligam na triptofano sintase, está o indol-3-
propanol fosfato (IPP) (WEYAND; SCHLICHTING, 1999) e cinco análogos do estado de
transição ariltioalquil fosfonados, que se ligam e inibem a subunidade α da triptofano
sintase (Figura 14) (SACHPATZIDIS et al., 1999). O IPP é um análogo de IGP que inibe a
subunidade α com um K
i
de 15 μM e foi o ponto inicial para a estratégia de desenho dos
inibidores ariltioalquil fosfonados (FINN et al., 1999). Estes inibidores foram desenhados
de forma a mimetizar o estado de transição formado durante a reação α da enzima e
possuem uma afinidade maior do que o substrato natural IGP e o análogo IPP. Estes
inibidores são derivados ariltioalquilfosfonados orto-substituídos que apresenta um átomo
de enxofre substituindo C2 (FINN et al., 1999; SACHPATZIDIS et al., 1999). Entretanto,
somente foram realizados estudos sobre o efeito inibitório dos inibidores ariltialquil
fosfonados sobre a triptofano sintase de Arabidopsis thaliana e S. typhimurium.
23
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figura 14. Fórmula molecular dos
ligantes de TRPS. A) inibidores
ariltioalquil fosfonados. 1) ácido 4-(2-
hidroxifeniltio)-1-butenilfosfônico, 2)
ácido 4-(2-
hidroxifeniltio)butenilfosfônico, 3) ácido
4-(2-aminofeniltio)-butilfosfônico, 4)
ácido 4-(2-hidroxi-5-fluorofeniltio)-
butilfosfônico e 5) ácido 4-(2-
hidroxifenilsulfinil)butilfosfônico.
B) IGP (Indol-3-glicerol fosfato e IPP
(Indol-3-propanol fosfato).
1
2
3
4
5
IPP
IGP
A B
1.6 InhA e o mecanismo de resistência ao antibiótico isoniazida
Apesar das enzimas comentadas anteriormente serem alvos em potencial para o
desenvolvimento de novos compostos que tenham atividade inibitória e, portanto,
possivelmente medicamentosas, nenhum composto foi desenvolvido especificadamente
para estas enzimas de Mycobacterium tuberculosis. Entretanto, uma das enzimas mais
estudadas para o desenvolvimento de drogas em M. tuberculosis, e que é alvo da
isoniazida, é a enzima dependente de NADH, 2-trans enoil-ACP (acyl carrier protein)
redutase (InhA). Esta enzima exibe alta especificidade para ácidos graxos de cadeia longa
(C
8
> C
16
) com grupo enoil tioester, que é consistente com o envolvimento desta enzima na
biossíntese de ácidos micólicos (QUÉRMAD et al. 1995). Ácidos micólicos, que por sua
24
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
vez, são ácidos graxos α-alquil-β-hidroxil, são os principais componentes da parede celular
de micobactérias (BRENNAN; NIKAIDO, 1995). Esta parede celular ou envelope celular
é uma característica única das micobactérias e apresenta um alto conteúdo de lipídios
constituindo uma barreira composta de ácidos micólicos ancorados a moléculas de
arabinogalactano, ligados ao peptídioglicano da membrana celular da bactéria (Figura 15)
(MOLLE et al., 2006). Essa parede é um grande obstáculo para a penetração de drogas na
célula da micobactéria, representando assim um grande desafio para o desenvolvimento de
novas drogas contra tuberculose.
Figura 15. Representação esquemática do envelope celular de micobactérias.
bicamada lipídica, peptídioglicano, arabinogalactano, micolato, acil
lipídeos,
lipoarabinomanano (LAM), porfirina
Dois tipos de sistemas de síntese de ácidos graxos (FAS) são conhecidos. FAS-I,
encontrado em vertebrados e está relacionado com a síntese de novo de ácidos graxos; e
25
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
FAS-II, que é usualmente encontrado em bactérias, e está relacionado com a elongação de
cadeias de ácidos graxos. Entretanto, M. tuberculosis apresenta ambos os sistemas, uma
cacterística que não é muito usual (DOVER et al., 2004). InhA é uma proteína pertencente
ao sistema de síntese micobacterial de ácidos graxos tipo II (FAS-II), que realiza a
elongação de precursores acil ácidos graxos, provindos do sistema de síntese micobacterial
de ácidos graxos tipo I (FAS-I) rendendo longas cadeias de carbono da parte meromicolato
dos ácidos micólicos (SCHROEDER et al., 2002). A função da InhA dentro do sistema
FAS-II é catalisar a redução dependente de NADH de duplas ligações trans entre a
posição C2 e C3 de substratos acil ácidos graxos (Figura 16) (QUEMARD et al., 1995).
Figura 16. Reação catalisada pela enzima InhA
A primeira estrutura da InhA em complexo com NADH foi determinada por Dessen
et al. (1995). A estrutura desta proteína apresenta uma estrutura terciária α/β com um
domínio apresentando um enovelamento Rossmann, que consiste em fitas β paralelas
ligadas por hélices α em uma ordem topológica β-α-β-α-β-α. A folha β formada apresenta
6 fitas na ordem 321456 (Figura 17). Este tipo de enovelamento classifica a InhA como
pertencente a superfamília das dehidrogenase/redutase de cadeia curta (SDR) (short
dehydrogenase/reductase) (DESSEN et al., 1995). Posteriormente à determinação desta
estrutura, outras estruturas da InhA com diferentes ligantes foram obtidas, entre elas os
complexos da InhA:NADH:isoniazida (ROSWARSKI et al., 1998), InhA:NADH:ácido
graxo C16 (ROSWARSKI et al., 1999) e o complexo da InhA com diversos ligantes
inibidores análogos de substrato ( KUO et al., 2003). Porém a conformação da InhA na
26
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
ausência de qualquer ligante ainda não foi elucidada, e assim, não se conhece a
conformação desta proteína antes da ligação da molécula de NADH.
Figura 17. Elementos de estrutura secundária da InhA de
Mycobacterium tuberculosis em complexo com NADH (em amarelo) e o
substrato C16 (em vermelho) (DESSEN et al., 1995)
A relação entre a ação da isoniazida e a inibição da InhA é um pouco complexa,
pois na verdade a isoniazida é uma pré-drogra que deve ser primeiramente convertida para
uma forma ativada através da enzima catalase-peroxidase (KatG) para um radical
isonicotínico acil (Figura 18) (ROZWARSKI et al., 1999). Este radical torna-se
covalentemente ligado ao anel nicotidamida do NADH ancorado no sítio ativo da InhA,
criando, desta forma, um NADH aducto que é o potente inibidor da InhA (ROZWARSKI
et al., 1998). A posição em que o radical isonicotínico acil liga-se à molécula de NADH é a
mesma região responsável pela transferência híbrida do átomo de hidrogênio que ocorre
durante a redução do substrato enoil-ACP (ROZWARSKI et al., 1998).
27
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
isoniazida
ADPR
Aducto Isoniazida-NADH
Figura 18. Mecanismo de ativação da isoniazida pela enzima
KatG. ADPR: ADP ribose.
Porém, tem sido observado mutações na estrutura do gene inhA que confere
resistência a isoniazida. Cepas apresentando mutações foram obtidas de isolados clínicos
de pacientes que vivem nos Estados Unidos (BASSO et al., 1998). Os mutantes isolados
apresentam a substituição de um único resíduo de aminoácido da estrutura primária da
InhA. Entre as mutações obtidas podemos destacar as seguintes: S94A, I21V e I47T. Estas
mutações estão bem caracterizadas cineticamente (RAWART; WHITTY; TONGE, 2003,
OLIVEIRA et al., 2006). Recentemente as estruturas cristalográficas dos complexos de
InhA:NADH com essas mutações foram determinadas (OLIVEIRA et al., 2006).
Entretanto, a estrutura da InhA mutante S94A em complexo com NAD-isoniazida foi
determinada por Rozwarski et al. (1998) a 2,8 Å de resolução, mostrando tanto o efeito
desta mutação sobre a estrutura como a influência na conformação dos resíduos do sítio
ativo (figura 19). Porém, os efeitos de outras mutações como a I21V e I47T sobre a
estrutura do complexo da InhA:NAD-isoniazida ainda não foram revelados. A resistência à
isoniazida por esses mutantes parece estar relacionada a uma baixa afinidade da proteína
pela molécula de NADH, que desta maneira, poderia promover a ligação da molécula de
acil substratos antes da molécula de NADH, desta forma dificultando a ligação da
molécula NAD-isoniazida, devido a um impedimento estérico. Lembrando que análise de
dados de cinética enzimática para InhA tem demonstrado que a seqüência de ligação do
28
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
NADH e ácidos graxos não é estritamente ordenada, mas ocorre uma preferência para o
NADH ligar-se primeiro à enzima (QUEMARD et al., 1995). Entretanto quando ocorre a
ligação do NAD-isoniazida primeiramente a baixa afinidade pela molécula de NADH
poderia permitir a rápida liberação desta molécula, e assim promovendo posterior ligação
da molécula de NADH permitindo a normal catálise da enzima (ROZWARSKI et al.,
1998, OLIVEIRA et al., 2006).
Figura 19. Contatos moleculares entre o complexo isonicotínico acil-NADH e o sítio ativo
da InhA. O grupo acil isonicotínico derivado da isoniazida está em vermelho, a porção
NADH está em azul, a cadeia lateral da InhA está em verde e a Ser94, o resíduo que causa
a resistência a isoniazida quando convertida em Ala está em rosa. Os números
representam a distância (em Å) entre os átomos selecionados (ROZWARSKI et al., 1998).
29
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
2. OBJETIVOS
Os objetivos do presente trabalho foram:
Cristalizar, coletar e processar dados de difração de raios X e resolver as estruturas
de duas enzimas da via do ácido chiquímico de M. tuberculosis, sendo elas a chiquimato
quinase complexada com ADP-ácido chiquímico (MtCQ-ADP-ác.Chiquímico), da
chiquimato quinase complexada com ADP e da corismato sintase na sua forma nativa
(MtCS).
Realizar a modelagem molecular da proteína triptofano sintase de M. tuberculosis
(MtTRPS) em complexo com inibidores indol propanol fosfato e uma série de 5 inibidores
ariltilalquil fosfonados.
Cristalizar, coletar e processar dados de difração de raios X e resolver as estruturas
das proteínas InhA selvagem e dos mutantes I21V e S94A em complexo com isoniazida e
a proteína mutante S94A na sua forma nativa.
30
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3. RESULTADOS
31
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.1 Crystallization and preliminary X-ray crystallographic analysis of chorismate
synthase from Mycobacterium tuberculosis.
Marcio V. B. Dias; Fernanda Ely; Fernanda Canduri; José H. Pereira; Jeverson Frazzon;
Luiz A. Basso; Mário S. Palma; Walter F. de Azevedo Jr; Diógenes S. Santos. Acta
Crystallographica Section D Biological Crystallography (ISSN 0907-4449), v. D60, p.
2003-2005, 2004.
Neste trabalho é descrita a cristalização e a análise preliminar de dados de difração
de raios X para a corismato sintase de M. tuberculosis (peso molecular de 42014 Da).
A proteína corismato sintase foi clonada, expressa e purificada pelo grupo de
Pesquisa do Prof. Dr. Diógenes Santiago Santos da PUC – Porto Alegre – RS. A proteína
inicialmente precipitada em sulfato de amônio foi dialisada contra tampão Tris HCl, 50
mM, pH 7.8. Para a cristalização foram utilizados os métodos de difusão de vapor, sistema
hanging drop e matriz esparsa. Após a obtenção dos primeiros cristais, a condição inicial
foi otimizada, variando-se a concentração de sal, precipitante e proteína. Foram obtidos
cristais para a corismato sintase solubilizada em Tris-HCl, 50 mM, pH 7,8, com dimensões
adequadas para a coleta de dados de difração de raios X, em uma condição composta por
Hepes-Na, 0,1 M, pH 7,5, Cloreto de Magnésio hexahidratado, 0,6 M e PEG 400, 25%. A
proteína estava em uma concentração de 60 mg.ml
-1
e a razão entre a solução de
cristalização e a solução de proteína era de 2:1. A coleta de dados de difração de raios X
foi realizada no LNLS. Os cristais obtidos são hexagonais, pertencente ao grupo espacial
P6
4
22 com dimensões aproximadas de 0,3x0,25x0,25 mm. O conjunto de dados obtido
para um cristal foi processado a 2.8 Å de resolução, apresentou uma completeza de 97,9%
e um R
sym
de 5,6%.
32
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
crystallization papers
Acta Cryst. (2004). D60, 2003±2005 DOI: 10.1107/S0907444904019869 2003
Acta Crystallographica Section D
Biological
Crystallography
ISSN 0907-4449
Crystallization and preliminary X-ray
crystallographic analysis of chorismate synthase
from Mycobacterium tuberculosis
Marcio Vinicius Bertacine Dias,
a
Fernanda Ely,
b
Fernanda
Canduri,
a,b
Jose
Â
Henrique
Pereira,
a
Jeverson Frazzon,
b
Luiz Augusto Basso,
b
Ma
Â
rio Se
Â
rgio Palma,
c
Walter Filgueira de
Azevedo Jr
a,b
* and
Dio
Â
genes Santiago Santos
d
*
a
Departamento de
Â
sica, UNESP, Sa
Ä
o Jose
Â
do
Rio Preto, SP 15054-000, Brazil,
b
Rede
Brasileira de Pesquisas em Tuberculose Grupo
de Microbiologia Molecular e Funcional,
Departamento de Biologia Molecular e
Biotecnologia, UFRGS, Porto Alegre,
RS 91501-970, Brazil,
c
Laboratory of Structural
Biology and Zoochemistry, CEIS/Department of
Biology, Institute of Biosciences, UNESP, Rio
Claro, SP 13506-900, Brazil, and
d
Centro de
Pesquisas em Biologia Molecular e Funcional/
Instituto de Pesquisas Biome
Â
dicas, Pontifõ
Â
cia
Universidade Cato
Â
lica do Rio Grande do Sul,
Porto Alegre, RS 90619-900, Brazil
Correspondence e-mail:
[email protected], diogenes@pucrs.br
# 2004 International Union of Crystallography
Printed in Denmark ± all rights reserved
The enzymes of the shikimate pathway are potential targets for the
development of new therapies because they are essential for bacteria
but absent from mammals. The last step in this pathway is performed
by chorismate synthase (CS), which catalyzes the conversion of
5-enolpyruvylshikimate-3-phosphate to chorismate. Optimization of
crystallization trials allowed the crystallization of homogeneous
recombinant CS from Mycobacterium tuberculosis (MtCS). The
crystals of MtCS belong to space group P6
4
22 (or P6
2
22) and diffract
to 2.8 A
Ê
resolution, with unit-cell parameters a = b = 129.7, c = 156.8 A
Ê
.
There are two molecules in the asymmetric unit. Molecular-
replacement trials were not sucessful. Heavy-atom derivative
screening is in progress.
Received 15 July 2004
Accepted 10 August 2004
1. Introduction
Tuberculosis is the second leading cause of
deaths worldwide, killing nearly 2 million
people each year. Most cases are in under-
developed countries; over the past decade,
tuberculosis incidence has increased in Africa,
mainly as a result of the burden of HIV
infection, and in the former Soviet Union,
owing to socioeconomic change and decline of
the health-care system (Frieden et al., 2003). In
1993, the gravity of the situation led the World
Health Organization (WHO) to declare
tuberculosis a global emergency in an attempt
to heighten public and political awareness
(Cole et al., 1998). Thus, newer and more ef®-
cient anti-tuberculosis drugs are needed.
Potential targets for the development of new
therapies are the enzymes of the shikimate
pathway, because they are essential for
bacteria, fungi and apicomplexan parasites, but
absent from mammals (Bentley, 1990; Roberts
et al., 1998). In microorganisms, the shikimate
pathway is used to synthesize the three
proteinogenic aromatic amino acids phenyl-
alanine (Phe), tyrosine (Tyr) and tryptophan
(Trp), the folate coenzymes benzoid and
naphtoid quinones and a broad range of mostly
aromatic secondary metabolites, including
siderophores (Dosselaere & Vanderleyden,
2001). This pathway consists of seven enzymes
that catalyse the sequential conversion of
erythrose-4-phosphate and phosphoenol
pyruvate to chorismate, the common precursor
of aromatic compounds (Herrmann & Weaver,
1999). The last step in this pathway is
performed by chorismate synthase, which
catalyzes the conversion of 5-enolpyruvyl-
shikimate 3-phosphate to chorismate via the
1,4-anti-elimination of phosphate and a proton,
a reaction that is unique in nature. The enzyme
has an absolute requirement for reduced FMN
as a cofactor, although the 1,4-anti-elimination
of phosphate and the C(6-pro-R) H atom does
not involve a net redox change. The role of the
reduced FMN in catalysis has long been
elusive. However, recent detailed kinetic and
bioorganic approaches have fundamentally
advanced our understanding of the mechanism
of action of chorismate synthase (Macheroux et
al., 1999).
Structural information is necessary for
structure-based design in drug discovery,
for a better understanding of the catalytic
mechanism and also for an understanding of
protein±ligand interaction (de Azevedo et al.,
1996, 1997, 2002). Although structures of CS
from some microrganisms have been deter-
mined (Ahn et al., 2004; Maclean & Ali, 2003),
the crystal structure of CS from Mycobacterim
tuberculosis has not been described. The
crystal structure of shikimate-pathway
enzymes will help in the development of new
drugs against tuberculosis and other infectious
diseases (de Azevedo et al., 2002; Pereira et al.,
2003). In this work, we initiated the structure
determination of the aroF-encoded MtCS,
which is composed of 401 residues with a
molecular weight of 41 792 Da. We report here
the crystallization and the preliminary X-ray
crystallographic study of MtCS.
2. Materials and methods
2.1. Cloning, protein expression and
purification
Synthetic oligonucleotide primers were
designed based on the aroF structural gene
sequence of M. tuberculosis H37Rv (Cole et al.,
1998), containing 5
H
NdeI and 3
H
BamHI
restriction sites. The PCR product was cloned
electronic reprint
33
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
crystallization papers
into pET-23a(+) (Novagen, USA) expres-
sion vector and transformed into Escher-
ichia coli Rosetta(DE3) host cells.
Transformed E. coli cells were grown in LB
medium at 310 K for 18 h in the absence of
IPTG, as used in previously reported
expression protocols (Oliveira et al., 2001;
Grossman et al., 1998), and harvested by
centrifugation. Although it is often argued
that the cost of IPTG limits the usefulness of
the lac promoter to high-added-value
products, it has previously been shown that
high levels of expression could be obtained
with pET vectors as cells entered the
stationary phase without the addition of
inducer in LB medium. It has been demon-
strated that when DE3 hosts are grown to
stationary phase in media lacking glucose,
cyclic AMP mediated derepression of both
the wild type and lacUV5 promoters occurs
(Oliveira et al., 2001). These authors also
proposed that cyclic AMP, acetate and low
pH are required for high-level expression in
the absence of IPTG induction when cells
approach the stationary phase in complex
media and that derepression of the lac
operon in the absence of IPTG may be part
of a general cellular response to nutrient
limitation.
The cell pellet was resuspended in 50 mM
Tris±HCl pH 7.8 (Sigma Co., USA), soni-
cated and centrifuged to remove cell debris.
The protein-puri®cation protocol included
the following steps: streptomycin sulfate
precipitation, ammonium sulfate precipita-
tion, Q Sepharose anion-exchange chroma-
tography, phenyl Sepharose hydrophobic
interaction chromatography and Mono Q
anion-exchange column (Amersham Phar-
macia Biotech, UK). A detailed description
of the cloning and expression of recom-
binant MtCS, the protein-puri®cation
protocol, N-terminal sequencing, mass
spectrometry and determination of the
oligomeric state of homogeneous M. tuber-
culosis chorismate synthase will be given
elsewhere (manuscript in preparation).
2.2. Crystallization
The puri®ed MtCS was concentrated and
dialyzed against 50 mM Tris±HCl buffer
pH 7.8 (Hampton Research, USA). The
®nal protein concentration was about
10 mg ml
À1
. Crystallization was performed
by the hanging-drop vapour-diffusion and
sparse-matrix methods (Jancarik & Kim,
1991) using tissue-culture multiwell plates
with covers (Linbro, ICN Biomedicals, Inc,
USA) at a temperature of 293 K. Each
hanging drop was prepared by mixing 1 ml
each of protein solution and reservoir solu-
tion and was placed over 700 ml reservoir
solution. Initial conditions were screened
using Crystal Screen I and II kits (Hampton
Research, USA).
Crystal optimization was carried out by
altering the concentration of the salt, preci-
pitant and protein and also the ratio
between the protein solution and reservoir
solution.
2.3. X-ray data collection
A data set was collected at a wavelength
of 1.427 A
Ê
using a synchrotron-radiation
source (Station PCr, LNLS, Campinas-
Brazil; Polikarpov et al., 1998). This wave-
length was previously set to optimize the
overall data-collection statistics. The data set
was collected from a single MtCS crystal
using a MAR CCD image-plate system. The
crystal was looped out from the drop and
¯ash-cooled. The PEG 400 present in the
crystallization conditions served as a cryo-
protectant. X-ray diffraction data were
collected at a temperature of 100 K under a
cold nitrogen stream generated and main-
tained with an Oxford Cryosystem. The
crystal was rotated through a total of 160
,
with a 1
oscillation range per frame, a
crystal-to-detector distance of 130 mm and
an exposure time of 60 s. Data were
processed on a Silicon Graphics Octane2
computer using the programs MOSFLM
(Leslie, 1990) and SCALA (Collaborative
Computational Project, Number 4, 1994).
2.4. Molecular-replacement trials
Molecular-replacement trials were carried
out with the program AMoRe (Navaza,
1994). The crystal structures of SpCs (PDB
code 1qxo; Maclean & Ali, 2003) and AaCs
(PDB code 1q1l; Viola et al., 2004) were
used as search models.
3. Results and discussion
The initial crystals were obtained with
reservoir solution comprising 0.2 M mag-
nesium chloride hexahydrate, 0.1 M Na
HEPES pH 7.5 and 30%(w/v) PEG 400. The
ratio of protein solution to well solution in
the drop was 1:1. After optimization of these
conditions, better crystals were obtained.
These diffracting crystals grew from a
reservoir solution containing 0.6 M mag-
nesium chloride hexahydrate, 0.1 M Na
HEPES pH 7.5 and 25%(w/v) PEG 400 and
the protein solution was concentrated to
60 mg ml
À1
in 50 mM Tris±HCl buffer pH
7.8. The ratio of protein solution to well
solution in the drop was 2:1. The crystals
grew reproducibly to approximate dimen-
sions of 0.3 Â 0.25 Â 0.25 mm within 2 d
(Fig. 1). The crystal diffracted to 2.8 A
Ê
with
relatively low mosaicity (0.35
). Fig. 2 shows
a typical X-ray diffraction pattern. A total of
92 610 measured re¯ections were merged
2004 Dias et al.
Chorismate synthase Acta Cryst. (2004). D60, 2003±2005
Figure 1
Hexagonal crystals of MtCS. Approximate dimen-
sions are 0.30 Â 0.25 Â 0.25 mm.
Figure 2
A typical diffraction pattern of the MtCS crystal with
1
oscillation range. The crystal diffracts to 2.8 A
Ê
resolution.
Table 1
Summary of data-collection statistics for MtCS.
Values in parentheses are for the highest resolution shell.
X-ray wavelength (A
Ê
) 1.427
Unit-cell parameters
a (A
Ê
) 129.74
b (A
Ê
) 129.74
c (A
Ê
) 156.77
Space group P6
4
22 or P6
2
22
No. measurements with I >2'(I) 92610
No. independent re¯ections 19341 (2716)
Completeness in the resolution
range 55.8±2.8 A
Ê
(%)
97.9 (97.9)
R
sym
² (%) 5.6 (16.5)
hI/'(I)i 3.8 (4.0)
Highest resolution shell (A
Ê
) 2.94±2.80
² R
sym
=
h
i
jIh
i
ÀhIhija
h
i
Ih
i
, where I(h)is
the intensity of re¯ection h,
h
is the sum over all re¯ections
and
h
is the sum over i measurements of re¯ection h.
electronic reprint
34
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
crystallization papers
into 19 341 unique re¯ections. The overall
R
sym
was 5.6% and the completeness was
97.9%. The crystal belongs to the hexagonal
space group P6
4
22 (or P6
2
22), with unit-cell
parameters a = b = 129.7, c = 156.8 A
Ê
.
Assuming the asymetric unit content to
be two monomers of molecular weight
41 792 Da, the V
M
value is 2.28 A
Ê
3
Da
À1
.
Assuming a value of 0.74 cm
3
g
À1
for the
protein partial speci®c volume, the calcu-
lated solvent content in the crystal is 45.96%
(Matthews, 1968). Table 1 summarizes the
data-collection statistics.
With the native data of MtCS to 2.8 A
Ê
resolution, molecular replacement was used
to attempt to solve the structure using
AMoRe (Navaza, 1994). Various search
models, including complete and modi®ed
structures of SpCS (PDB code 1qxo) and
AaCS (1q1l), did not yield any meaningful
results. This was most likely to be because of
conformational differences between the
search models and the structure under study.
Heavy-atom screening is in progress.
This work was supported by grants from
FAPESP (SMOLBNet, Proc. 01/07532-0, 02/
04383-7, 04/00217-0), CNPq, CAPES and
Instituto do Mile
Ã
nio (CNPq-MCT) to DSS
and LAB. WFA (CNPq, 300851/98-7), MSP
(CNPq, 500079/90-0), DSS (CNPq, 304051/
1975-06), LAB (CNPq, 520182/99-5) and JF
(CNPq, 301131/2003-01) are researchers for
the National Research Council.
References
Ahn, H. J., Yoon, H. J., Lee, B. & Suh, S. W. (2004).
J. Mol. Biol. 27, 903±915.
Azevedo, W. F. de Jr, Canduri, F. & da Silveira
N. J. F. (2002). Biochem. Biophys. Res.
Commun. 293, 566±571.
Azevedo, W. F. de Jr, de Oliveira, J. S., Basso, L. A.,
Palma, M. S., Pereira, J. H., Canduri, F. &
Santos, D. S. (2002). Biochem. Biophys. Res.
Commun. 295, 142±148.
Azevedo, W. F. de Jr, Leclerc, S., Meijer, L.,
Havlicek, L., Strnad, M. & Kim, S.-H. (1997).
Eur. J. Biochem. 243, 518±526.
Azevedo, W. F. de Jr, Mueller-Dieckmann, H. J.,
Schulze-Gahmen, U., Worland, P. J., Sausville,
E. & Kim, S.-H. (1996). Proc. Natl Acad. Sci.
USA, 93, 2735±2740.
Bentley, R. (1990). Crit. Rev. Biochem. Mol. Biol.
25, 307±384.
Cole, S. T. et al. (1998). Nature (London), 393,
537±544.
Collaborative Computational Project, Number 4
(1994). Acta Cryst. D50, 760±763.
Dosselaere, F. & Vanderleyden, J. (2001). Crit.
Rev. Microbiol. 27, 75±131.
Frieden, T. R., Sterling, T. R., Munsiff, S. S., Watt,
C. J. & Dye, C. (2003). Lancet, 362, 887±899.
Grossman, T. H., Kawasaki, E. S., Punreddy, S. R.
& Osburne, M. S. (1998). Gene, 209, 95±103.
Herrmann, K. M. & Weaver, L. M. (1999). Annu.
Rev. Plant. Physiol. Plant. Mol. Biol. 50, 473±
503.
Jancarik, J. & Kim, S.-H. (1991). J. Appl. Cryst. 24,
409±411.
Leslie, A. G. W. (1990). Crystallographic
Computing, edited by D. Moras, A. D. Podjarny
& J. C. Thierry, pp. 27±38. Oxford University
Press.
Macheroux, P., Schmid, J., Amrhein, N. & Schaller,
A. (1999). Planta, 207, 325±334.
Maclean, J. & Ali, S. (2003). Structure, 11, 1499±
1511.
Matthews, B. W. (1968). J. Mol. Biol 33, 491±
497.
Navaza, J. (1994). Acta Cryst. A50, 157±163.
Oliveira, J. S., Pinto, C. A., Basso, L. A. & Santos,
D. S. (2001). Protein Expr. Purif. 22 , 430±435.
Pereira, J. H., Canduri, F., de Oliveira, J. S., da
Silveira, N. J., Basso, L. A., Palma, M. S., de
Azevedo, W. F. Jr & Santos, D. S. (2003).
Biochem. Biophys. Res. Commun. 312, 608±
614.
Polikarpov, I., Perles, L. A., de Oliveira, R. T.,
Oliva, G., Castellano, E. E., Garratt, R. C. &
Craievich, A. (1998). J. Synchrotron Rad. 5, 72±
76.
Roberts, F., Roberts, C. W., Johnson, J. J., Kyle,
D. E., Krell, T., Coggins, J. R., Coombs, G. H.,
Milhous, W. K., Tzipori, S., Ferguson, D. J.,
Chakrabarti, D. & McLeod, R. (1998). Nature
(London), 395, 801±805.
Viola, C. M., Saridakis, V. & Christendat, D.
(2004). Proteins, 54, 166±169.
Acta Cryst. (2004). D60, 2003±2005 Dias et al.
Chorismate synthase 2005
electronic reprint
35
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.2 Structure of chorismate synthase from Mycobacterium tuberculosis.
Marcio V.B. Dias; Júlio C. Borges; Fernanda Ely; José H. Pereira; Fernanda Canduri;
Carlos H.I. Ramos; Jeverson Frazzon; Mário S. Palma; Luis A. Basso; Diógenes S. Santos;
Walter F. de Azevedo Jr. Journal of Structure Biology (ISSN 1047-8477), v. 154, p. 130-
143, 2006.
Neste trabalho é descrita a estrutura cristalográfica da proteína corismato sintase de
M. tuberculosis (peso molecular de 42014 Da).
A estrutura da corismato sintase de M. tuberculosis foi resolvida a 2,65 Å de
resolução através da técnica de substituição molecular utilizando como modelo de busca a
estrutura da corismato sintase de S. pneumoniae (MACLEAN; ALI, 2003). Os resíduos
não conservados entre as duas estruturas foram trocados para alanina, com exceção de
glicinas. A função translação obtida na substituição molecular apresentava um coeficiente
de correlação de 37,6% e um R-factor de 53%. Após a confirmação do grupo espacial
correto (P6
4
22), realizou-se refinamento cristalográfico. O modelo para a corismato sintase
de M. tuberculosis foi iterativamente ajustado com base em mapas de densidade eletrônica
com coeficientes 2Fo – 2Fc e Fo – Fc. A estrutura final obtida apresenta um R-factor de
16,2% e um R-free de 22,1% e uma boa qualidade estereoquímica, apresentando 99,3%
dos resíduos em regiões favoráveis do gráfico de Ramachandran e apenas 0,3% em regiões
não permitidas.
A estrutura da corismato sintase de M. tuberculosis é similar às outras estruturas
determinadas anteriormente para outros microorganismos. A unidade assimétrica do cristal
apresenta um único monômero, que pertence à classe de proteínas α/β. Este monômero
apresenta somente um domínio, cuja topologia estrutural é um sanduíche β-α-β. Este
sanduíche apresenta um centro formado de hélices que se encontra entre duas camadas de
36
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
folhas β formadas por quatro fitas. Este domínio é rodeado por loops e trechos discretos de
hélices α e fitas β. Entretanto, a estrutura cristalográfica da corismato sintase de M.
tuberculosis pode ser descrita como um tetrâmero formado por dímeros de dímeros, com
simetria 222, sendo que a principal característica da interface de dimerização é uma folha β
antiparalela intermonomérica, formada por oito fitas, sendo quatro de um monômero e
quatro do outro monômero adjacente. A comparação do sítio ativo da corismato sintase de
M. tuberculosis com a de outros organismos mostra que muitos resíduos são conservados
entre estas estruturas. Entretanto a posição da cadeia lateral de vários resíduos apresenta-se
em conformações diferentes, talvez devido à ausência de ligantes em nossa estrutura e a
presença de ligantes nas outras estruturas que foram utilizadas para a análise. Portanto,
devido à ausência de ligantes em nossa estrutura e a determinação da posição de quase
todos os resíduos do sítio ativo, com exceção dos resíduos 49-52, pode-se inferir
importantes movimentos ocorridos pela ligação do substrato EPSP e da co-enzima FMN no
sítio ativo desta proteína.
37
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Structure of chorismate synthase from Mycobacterium tuberculosis
Marcio V.B. Dias
a
,Ju
´
lio C. Borges
a
, Fernanda Ely
b
, Jose
´
H. Pereira
a
,
Fernanda Canduri
a,c
, Carlos H.I. Ramos
d
, Jeverson Frazzon
b
,Ma
´
rio S. Palma
e
,
Luis A. Basso
f
, Dio
´
genes S. Santos
f,
*
, Walter F. de Azevedo Jr.
a,g,
*
a
Programa de Po
´
s-Graduac¸a
˜
o em Biofı
´
sica Molecular, Departamento de
´
sica, UNESP, Sa
˜
o Jose
´
do Rio Preto, SP 15054-000, Brazil
b
Rede Brasileira de Pesquisa em Tuberculose, Grupo de Microbiologia Molecular e Funcional, Departamento de Biologia Molecular e Biotecnologia,
UFRGS, Porto Alegre, RS 91501-970, Brazil
c
Departamento de Morfofisiologia, Laborato
´
rio de Bioquı
´mica,
CCBS, Universidade Federal de Mato Grosso do Sul,
Campo Grande, MS, CEP 79070-900, Brazil
d
Centro de Biologia Molecular Estrutural, Laborato
´
rio Nacional de Luz
´nc
rotron, Campinas SP, Brazil
e
Departamento de Biologia, CEIS/IBRC, UNESP, Rio Claro, SP, CEP 13506-900, Brazil
f
Pontifı
´
cia Universidade Cato
´
lica do Rio Grande do Sul, Centro de Pesquisa em Biologia Molecular e Funcional, Porto Alegre, RS, Brazil
g
Faculdade de Biocie
ˆ
ncias—Pontifı
´
cia Universidade Cato
´
lica do Rio Grande do Sul, Av. Ipiranga, 6681. Porto Alegre-RS CEP 90619-900, Brazil
Received 17 July 2005; received in revised form 8 November 2005; accepted 9 December 2005
Available online 17 January 2006
Abstract
In bacteria, fungi, plants, and apicomplexan parasites, the aromatics compounds, such as aromatics amino acids, are synthesized
through seven enzymes from the shikimate pathway, which are absent in mammals. The absence of this pathway in mammals make
them potential targets for development of new therapy against infectious diseases, such as tuberculosis, which is the world’s second
commonest cause of death from infectious disease. The last enzyme of shikimate pathway is the chorismate synthase (CS), which is
responsible for conversion of the 5-enolpyruvylshikimate-3-phosphate to chorismate. Here, we report the crystallographic structure
of CS from Mycobacterium tuberculosis (MtCS) at 2.65 A
˚
´
resolution. The MtCS structure is similar to other CS structures, present-
ing bab sandwich structural topology, in which each monomer of MtCS consists of a central helical core. The MtCS can be
described as a tetramer formed by a dimer of dimers. However, analytical ultracentrifugation studies suggest the MtCS is a dimer
with a more asymmetric shape than observed on the crystallographic dimer and the existence of a low equilibrium between dimer
and tetramer. Our results suggest that the MtCS oligomerization is concentration dependent and some conformational changes must
be involved on that event.
Ó 2005 Elsevier Inc. All rights reserved.
Keywords: Chorismate synthase; Crystallography; Analytical ultracentrifugation; Mycobacterium tuberculosis; Shikimate pathway
1. Introduction
In mammals, diet has to provide the essential amino
acids, such as, phenylalanine, tryptophan, and tyrosine.
In bacteria, fungi, plants, and apicomplexan parasites,
these aromatic amino acids are synthesized through the
complex shikimate pathway (Bentley, 1990; Macheroux
et al., 1999). Furthermore, the shikimate pathway provides
these organisms the basic building blocks for the synthesis
of the other aromatic compounds required for different
functions as UV protection, electron transport, signaling,
iron uptake, etc. The shikimate pathway is responsible
for conversion of
D-erythrose-4-phosphate and phospho-
enolpyruvate to shikimate and subsequently to the dihyd-
roaromatic compound chorismate (Bentley, 1990; Dossela-
ere and Vanderleyden, 2001). The absence of shikimate
pathway in mammals has rendered its enzymes as
www.elsevier.com/locate/yjsbi
Journal of Structural Biology 154 (2006) 130–143
Journal of
Structural
Biology
1047-8477/$ - see front matter Ó 2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.jsb.2005.12.008
*
Corresponding authors. Fax: +55 17 51 3220 3629.
E-mail addresses: diogenes@pucrs.br (D.S. Santos), walter.junior@
pucrs.br (W.F. de Azevedo Jr.).
38
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
potentially targets for the development of new therapy
against infectious disease, such as tuberculosis (TB).
1
TB is considerate one of the most serious global public
health challenge of the 21st century. It is the world’s second
commonest cause of death from infectious disease, after
acquired immune deficiency syndrome. According to a
recent report compiled by the World Health Organization,
the total number of new cases of TB worldwide in 2002 ha s
increased to around 9 million (Duncan, 2004; WHO, 2004),
from which 2–3 million people died, despite the chemother-
apy available (WHO, 2004). Thes e deaths are mostly of
young adults but also include about 100 000 children unde r
the age of 5 years (Gandy and Zumla, 2002). The key driver
of the increase of TB is its synergy with the human immu-
nodeficiency virus epidemic, which has a devastating
impact in some parts of the world, such as, the African
region (Duncan, 2004). Thus, there is an urgent need for
new anti-mycobacterial inhibitors.
The enzymes of shikimate pathway are good candidates
for development of new therapies against TB. Enzymes
from this metabolic pathway have been submi tted to inten-
sive structural studies (Azevedo et al., 2002; Basso et al.,
2005; Pereira et al., 2003, 2004; Silveira et al., 2005). The
last enzyme from this pathway is the ch orismate synthase
(CS), which catalyzes the conversion of the 5-enolpyruvyls-
hikimate-3-phosphate (EPSP) to chorismate. The CS reac-
tion comprises an anti-1,4-elimination of the 3-phosphate
group and the C(6proR)-hydrogen (Bornemann et al.,
2003; Hill and Newkome, 1969; Macheroux et al., 1998).
It is the only enzymatic reaction known of such transfor-
mation in biological systems, making the CS a unique
enzyme in the nature. The CS requires reduced flavin
mononucleotide (FMN), an essential cofactor typically
found in many biologi cal redox reactions. Surprisingly,
the reaction catalyzed by CS does not involve an overall
change in redox state (Bornemann et al., 1996;Kitzing
et al., 2004; Macheroux et al., 1996; Macheroux et al.,
1999). According Bornemann et al. (1995), the reduced
FMN donates an electron to EPSP to facilitate the loss
of the phosphate and receive it back after the reaction.
So, only flavin in its reduced form is functional and it is
not consumed during the reaction (Macheroux et al.,
1998; Welch et al., 1974). Furthermore, two classes of CS
are distinguished among the microorganisms that possess
shikimate pathway. CS from yeasts ha ve the ability to
use b-nicotinamide adenine dinucleotide phosphate
(NAD(H)P) for the reduction of oxidized FMN, having
therefore, an additional catalytic activity and wher efore
are called bifunctionals, while all the other CS lacking it
are called monofunctional (Macheroux et al., 1999).
There are crystal structures available of CS from three
bacteria: CS from Streptococcus pneumoniae complex ed
with FMN and EPSP (Maclean and Ali, 2003); CS from
Helicobacter pylori co mplexed with FMN (Ahn et al.,
2004; and native CS from Aquiflex aelicus (Viola et al.,
2004) (PDB access codes: 1QXO, 1UMO, and 1Q1L,
respectively) and one CS from yeast (native CS of
Saccharomices cerevisae) (PDB access code: 1R53,
Quevillon-Cheruel et al., 2004). In all these structures, the
CS is presented as tetramer composed of two dimers, like
a dimer of dimers. The monomer comprises a single large
core domain, which is surrounded by loops and discrete
stretches of a-helix and b-sheet. Here, we present structural
data of the CS from M. tuberculosis (MtCS) in its native
form. The structure of the MtCS was determined at
2.65 A
˚
resolution and it presents a bab architecture.
The analytical ultracentrifugation data suggest that the
MtCS is predominantly a dimer in solution presenting
equilibrium with a tetrameric form. This structure can help
us understand the action mecha nism of the MtCS.
2. Experimental procedures
2.1. MtCS recombinant production, purification, and
crystallization
The MtCS cloning, expression, purification, and crystal-
lization were reported elsewhere (Dias et al., 2004). The
protein concentration was determined spectrophotometri-
cally as described by Edelhock (1967), using a calculated
extinction coefficient for denatured proteins (Gill and von
Hippel, 1989).
2.2. Circular dichroism spectroscopy
Circular dichroism (CD) measurements were performed
using a Jasco J-810 spectropolarimeter with the tempera-
ture controlled by a Peltier-type control system PFD
425S. The data were collected at a scanning rate of
100 nm/min with a spectral band width of 1 nm and using
a 1 mm path length cell. The thermal-induced unfolding
experiments followed by CD were performed at a scan rate
of 1 °C/min. The average of three unfolding curves was
used to build the MtCS thermal-unfold profile and the tem-
perature at the midpoint of the unfolding transition was
determined by fitting with Gaussians of the first derivative
function.
2.3. Structure determination
The data set of MtCS was collected at a wavelength of
1.427 A
˚
using the Synchrotron Radiation Source (Station
PCr, LNLS, Campinas—Brazil) (Polikarpov et al., 1998)
and a CCD detect or (MARCCD). The crystal was flash-
frozen at 104 K under cold nitrogen stream generated
and maintained with an Oxford Cryosystem. The data set
was processed up to 2.65 A
˚
resolution using the program
1
Abbreviations used: TB, tuberculosis; CS, chorismate synthase; MtCS,
chorismate synthase from Mycobacterium tuberculosis; EPSP, 5-eno-
lpyruvylshikimate-3-phosphate; FMN, flavin mononucleotide; CD, circu-
lar dichroism; RMSD, root-mean-square deviation; AUC, analytical
ultracentrifugation.
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 131
39
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
MOSFLM ( Leslie, 1992) and scaled with SCALA (CCP4,
1994).
The crystal structure of the MtCS was determined by
standard molecular replac ement methods implemented in
the program AMoRe (Navaza, 2001), using as search mod-
el the monomer structure of Streptococcus pneumoniae CS
(SpCS) (PDB access code: 1QXO) (Maclean and Ali,
2003), in which nonconserved residues were changed to
alanines. The rotation search was calculated using data
in the range of 12.0–3.0 A
˚
resolution, which gave a solu-
tion with a correlation of 7.3%. The translation search
yielded a unique solution with a correlation of 37.6%
and an R-factor of 53.0% was obtained. At this stage,
the correct space group was confirmed to be P6
4
22 by visu-
alization of crystal packing using the program O (Jones
et al., 1991). Further crystallographic refinement was car-
ried out using X-PLOR and simulated annealing (Brunger,
1992). The protein model obtained by molecular replace-
ment was refined resulting R-factor and R-free of about
33.6 and 43.1%, respectively. Throughout the refinement,
the composite electron-density maps with coefficients
2F
o
À F
c
and F
o
À F
c
were calculated with X-PLOR and
visualized using XtalView/Xfit (McRee, 1999), and the
model was built and iteratively adjusted. After several
refinement steps the R-factor and the R-free decreased to
25.7 and 32.5%, respectively. Further, the group B factors
were adjusted and it was used the data from 8.0 to 2.65 A
˚
.
The model was subjected to further refinement using the
maximum-likelihood based program REFMAC5 (Murs-
hudov et al., 1997) using data between 52.93 and 2.65 A
˚
´
resolution. At this stage, water molecules were added to
the model using the program XtalView and the ARP rou-
tine of REFMAC5. Refinement converged to R-factor and
R-free values of 16.2 and 22.1%, respectively. The correct-
ness of the stereochemi stry of the model was checked using
PROCHECK (Laskowski et al., 1994). Root-mean-square
deviation (RMSD) differences from ideal geometries for
bond lengths, angles, and dihedrals were calculated with
X-PLOR (Bru
¨
nger, 1992). Atomic models were superposed
using the program LSQKAB from CCP4 (CCP4, 1994).
The molecular surface areas have been calculated using
the program AREAIMOL/RESAREA (CCP4, 1994).
The PARMODEL (Ucho
ˆ
a, 2004) was used in the analysis
of the final model.
2.4. Gel filtration
Gel filtration was performed using a Superdex S-200
column (Amershan Pharmacia Biotech) in FPLC system.
The column was equilibrated in 50 mmol/L of Tris–HCl
buffer, pH 7.8, containing 200 mmol/L of NaCl. Protein
standards used were ribonuclease A (13.7 kDa), chymo-
trypsinogen (25 kDa), ovalbumin (43 kDa), albumin
(67 kDa), aldalose (158 kDa), catalase (232 kDa), ferritin
(440 kDa), and thyroglobulin (669 kDa) (Amershan Phar-
macia Biotech). Elution time of MtCS was recorded and
molecular weight was calculated by estimating the elution
volumes of standards of known molecular weight. The
MtCS were loaded on the gel filtration column at a con-
centration of 1 mg/mL.
2.5. Analytical ultracentrifugation
Sedimentation velocity and sedimentation equilibrium
experiments were performed using a Beckman Optima
XL-A analytical ultracentrifuge. The sedimentation veloc-
ity experiments were carried out at 20 °C and 25000 rpm
(AN-60Ti rotor) with the scan data acquisition at 232
and 236 nm. MtCS was tested in concentrations of
300–1000 lg/mL in 20 mmol/L Tris–HCl buffer, pH 8.0,
containing 50 mmol/L NaCl and 1 mmol/L b-mercap-
toethanol. The analysis involved fitting a model of absor-
bance versus cell radius data by nonlinear regression. The
analysis was performed with the ORIGIN software
package (MicroCal Software) supplied with the instru-
ment. The second moment (Goldberg, 1953) an d the
sedimentation time derivative (g (s*) integral distribution)
(Stafford, 1994) methods were used to analyze the sedi-
mentation velocity experiments. These methods allow
the calculation of the apparent sedimentation coefficient
(s*), the diffusion coefficient D, and the molecular mass
MM. The ratio of the sedimentation coefficient to diffu-
sion coefficient gives the molecular mass by use of the
follow equation:
MM ¼
sRT
Dð1 À VbarqÞ
; ð1Þ
where R is the gas constant and T is the absolute tempera-
ture. The software Sednterp (www.jphilo.mailway.com/
download.htm) was used to estimate protein partial specific
volume at 20 °C(Vbar = 0.7329 mL/g), buffer density
(q = 1.00087 g/mL), and viscosity (g = 1.0126 · 10
À2
P),
and the s
max
for globular proteins of about 41.8, 83.6,
and 167.2 kDa. This software was also used to estimate
the standard sedimentation coefficients (s
20,w
) for each pro-
tein concentration that were used to estimate the s
20,w
at
0 mg/mL protein concentration by extrapolation (s
0
20;w
Þ.
That procedure minimizes interferences caused by temper-
ature, viscosity solution, and molecular crowd (Laue,
2001).
The sedimentation equilibrium experiments were made
at 20 °C at speeds of 8000, 10000, and 12 000 rpm with
the AN-60Ti rotor, scan data acquisition at 236 nm and
at protein concentration of 300, 450, and 600 lg/mL. Brief-
ly, the samples were accelerated to 8000 rpm and scans
were collected until no changes in the scan were observed.
Three scans were obtained at 8000 rpm. Similar procedures
were made at 10000 and 12 000 rpm. The self-association
method was used to analyze the sedimentation equilibrium
experiments using several models of association to fit the
MtCS data. The distribution of the protein along the cell,
obtained in the equilibrium sedimentation experiments,
was fitted with the following equation (Johnson et al.,
1981):
132 M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143
40
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
C ¼ C
0
e
M ð1 À VbarqÞx
2
ðr
2
À r
0
Þ
2RT

; ð2Þ
where C is the protein concentration at radial position r , C
0
is the protein concentration at radial position r
0
, and x is
the centrifugal angular velocity.
The HydroPro software (Garcı
´
a de la Torre et al., 2000)
was applied to estimate the standard sedimentation coeffi-
cient s
20,w
, starting from the high resolution MtCS struc -
ture: as a monomer, dimer an d tetramer. The HydroPro
software was setup with the radius of the atomic elements
of 2.5 A
˚
´
, with sigma factors from 5 to 8 (as indicated by
supplier) and minibeads radius (SIGMIN and SIGMAX)
from 6 to 2 A
˚
´
after initial evaluation of the two extremes.
The parameters MtCS Vbar, and q and g (for standard
conditions) were estimated using the software Sednterp as
described above.
3. Results and discussion
3.1. MtCS spectropolarimetry analysis
The observed spectrum of CD data is shown in Fig. 1,
and it is similar to the one obtained for CS from Escherich-
ia coli (EcCS) (Macheroux et al., 1998). The deconvolution
of CD data predicted that the MtCS has nearly 30 and 18%
of a-helices and b-sheet, respectively. This result is in agree-
ment with the crystallographic structure presented here (34
and 17% of a-helices and b-sheet, respectively). the MtCS
thermal stability is presented in Fig. 1 (inset). The data
show that the MtCS has two well-characterized thermal
unfolding transitions: one at 51.5 °C and another at
64 °C, which is disagreement with the data from thermal
unfolding of the EcCS. The data for EcCS suggests only
one unfolding transition at 54 °C(Fitzpatrick et al.,
2001), despite of the experimental conditions were different
than that applied for the MtCS. Thus, the data obtained
here, may suggest that the MtCS possesses two well-defined
domains, each one presenting different thermal stabilities,
despite MtCS crystal structure does not show the existence
of these domains (see below). Then, other hypotheses is
that the protein may loss its oligomeric status before under-
going the fully unfold. Furthermore, as the protein is a
dimer in solution (see below), the secondary structures on
the flanks of the dimerization interface core can also under-
go unfolded at 51.5 °C and the core lost its secondary
structure at 64 °C. The MtCS unfolding mechanism is cur-
rently under investigation.
3.2. MtCS presents a ba
b architecture
The crystal structure of the MtCS was solved at 2.65 A
˚
using the Synchrotron Radiation Source (Polikarpov et al.,
1998), and the structure was determined by molecular
replacement. The final statistics for data processing and
refinement are in Table 1. There is one molec ule of MtCS
in the asymm etric unit, containing residues 1–48 and 53–
392 and 243 water molecules (Fig. 2A). The residues 49–
52 and 393–401 did not present electronic density, and they
were omitted in the final model. The analysis of stereo-
chemical quality, G-factor and 3D-profile are presented in
Table 1.
The structure of MtCS is similar the other CS structures
deposited in the Protein Data Bank (PDB) (PDB access
code: 1QXO, 1Q1L, 1UMO, and 1R53; Ahn et al., 2004;
Maclean and Ali, 2003; Quevillon-Cheruel et al., 2004; Vio-
la et al., 2004, respectively). The structure of the MtCS
monomer belongs to the a/b class. Each monomer of
MtCS contains 13 a-helices and 17 b-sheets, which fold
into an approximate dimension of 75 A
˚
· 47 A
˚
· 36 A
˚
.
MtCS presents only one domain and its dominant structur-
al topology is a bab sandwich, in which each monomer
of MtCS contains a central helical core (formed by helices
a1, a5, a11, and a8). This helical core is sandwiched
between two four-stranded antiparallel b-sheets (b1, b2,
b7, b4, and b8, b9, b15, b10), which are surrounded by
loops and discrete stretches of a-helix and b-sheet. The
topology of MtCS is shown in
Fig. 2B and can be observed
that the core of structure of MtCS has pseudo 2-fold sym-
metry, despite the molecule does not present such
symmetry.
The crystal structure obtained for MtCS may be
described as a tetramer formed by a dimer of dimers, pre-
senting approximately 222 symmetry. The intermolecular
interactions among the monomers A and D form the first
dimer and those among the monomers B and C forming
the second dimer. But each monomer of CS is in contact
with three others, creating an intricate packing arrange-
ment (Fig. 3). The contact areas between the monomers
are 796.0, 1944.0, and 3977.0 A
˚
2
for the unit pairs AB/
CD, AC/BD, and AD/BC, respectively, and many residues
involved on the interface of tetramerization are conserved
062052042032022012002
51-
01-
5-
0
5
10
15
-12
-10
-8
-6
-4
-2
80706050403020
( erutarepmeT
o
)C
[] (Deg.cm
2
.dmol
-1
)x10
3
)mn( htgnelevaW
[θ]
222nm
(Deg.cm
2
.dmol
-1
)x10
3
Fig. 1. Circular dichroism analyses. Residual molar ellipticity of the
MtCS was measured from 195 to 260 nm in 20 mmol/L Tris–HCl buffer
(pH 8.0) containing 50 mmol/L of NaCl, and 1 mmol/L b-mercap-
toethanol at 20 °C. The amount of secondary structure was estimated by
CDNN Deconvolution software (Bo
¨
hm et al., 1992). Thermal stability of
MtCS was analyzed by CD at 222 nm.
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 133
41
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
in several sequences from bacteria (Fig. 4). Fig. 5 introduc-
es the interface contact area between the monomers pairs
AB, AC, and AD, emphasizing the molecular surface.
The dimers form ed with monomers A/D or B/C of CS
are charact erized by a large antiparallel eight-stranded b-
sheet that is formed by combination of two sheets, one
from each monomer (Fig. 5A). This large b-sheet is the
major feature of the MtCS dimer interface and may be said
that the two strands b10 are the principal region for the
dimer stabilization. Another secondary structure element
involved in stabilization of the dimer is the helix a8 that
interacts with a8 from another monomer. These interac-
tions contribute highly to dimer stabi lization, presenting
considerable amount of hydrophobic interactions on the
surface at the dimer interface. The dimer also presents
two additional four-strand antiparallel b-sheet, one of each
side of the core formed by association between the b11 and
b12 from one monomer and b16 and b17 of another mono-
mer. Moreover, the dimer interface presents several inter-
actions (44 hydrogen bonds), mainly involv ing loops and
the a3 and a4.
The main residues involved in the MtCS dimerization
are Lys123, Tyr124, Arg135, Gly171, Asp217, Gly220,
Gly233, Gly235, Asp245, Ile25 5, Lys259, Gly274, and
Pro316, which are conserved in almost all CS sequences
from bacteria including the four-solved structures (Fig. 4,
see below). These results suggest that probably all CS are
at least dimeric due the high number of hydrogen bonds
and conservation of the residues involved in this interac-
tion. Fig. 4 shows the alignment of 16 sequences of CS
from bacteria evidencing the conserved residues involved
on the interface of two monomers A and D or B and C,
which form the dimer.
The MtCS crystal structure clearly shows that the inter-
face interactions between the dimer AD and BC, which
form of tetrameric structure of CS involve both hydropho-
bic and polar residues, and some regions present comple-
mentary of charge.
On the other hand, the forces that stabilize the interac-
tion between the monomers A and B are made mainly by
two salt-bridges and six hydrogen bonds localized in the
following regions: b2, a7, a11, a12, and L29, mainly
between a12 from one monomer with the adjacent from
other monomer. The number of the interactions between
the monomers A and C is higher than those observed for
monomers A and B. The interaction between monomer A
and C is made mainly by 18 hydrogen bonds localized prin-
cipally in regions of loops. The protein regions responsible
for these interactions are: L1, a1, L5, b5, L6, b7, L8, L9,
a3, L10, and L29. The predominant region of interaction
between monomer A and C is L8, which interact with the
L8 from the adjacent monomer.
3.3. Comparison between MtCS structure and others CS
structures
The monomeric structure of MtCS was compared with
five other CS structures from four organisms: three from
bacteria (Aquifex aeolicusAaCS; Streptococcus pneumo-
niaeSpCS and Helicobacter pyloriHpCs) and one from
yeast (Saccharomyces cerevisiaeScCS) deposited in the
PDB (PDB access code: 1Q1L, 1QXO, 1UMO, and
1R53, respectively) (Ahn et al., 2004; Maclean and Ali,
2003; Quevillon-Cheruel et al., 2004; Viola et al., 2004,
respectively). Maclean and Ali (2003) reports the SpCS
structure (which shares with MtCS, on their amino acid
Table 1
Crystallographic data, refinement statistics, and analysis of the quality
from MtCS
1° Crystal data
Unit-cell parameters a = b = 129.74, c = 159.77
Space group P6
4
22
Number of measurements with I > 2 r (I) 92610
Number of independent reflections 19341
Completeness (%) (outermost shell) 97.90 (97.9)
R
sym
a
(%) (outermost shell) 5.40 (16.50)
Highest resolution shell (A
˚
) 2.94–2.65
2° Refinement data
Resolution range (A
˚
) 52.93–2.65
Reflections used for refinement 20093
Number of atoms 3083
Number of residues 388
Number of O atoms 243
Final R-factor (%)
b
16.2
Final R-free final (%)
c
22.1
Correlation coefficient F
o
À F
c
(%) 95.6
B values (A
˚
2
)
d
Main chain 33.81
Side chain 36.05
3° Quality of structure
Observed RMSD from ideal geometryl
Bonds lengths (A
˚
) 0.021
Bonds angles (°) 2.655
G-factor
e
Torsion angles À0.49
Covalent geometry À0.15
Overall À0.27
3D profile
f
(S) = 170.39; IS = 177.45;
S/IS = 0.96 IS
Ramachandran plot (%)
Favorable 92.2
Additional allowed 7.5
Generously allowed 0.0
Disallowed 0.3
a
R
sym
= 100 PjI (h)_ÆI (hj/PI(h) with I (h), observed intensity and
ÆI (h)æ, mean intensity of reflection h overall measurement of I (h).
b
R-factor = 100 ·
P
(jF
obs
À F
calc
j)/
P
(F
obs
), the sums being taken over
all reflections with F/r(F)>2r ( F).
c
R-free = R-factor for 10% of the data that were not included during
crystallographic refinement.
d
B values = average B values for all non-H atoms.
e
Ideally, scores should be above À0.5. Values below À1.0 may need
investigation.
f
Total score is the sum of the 3D–1D scores (statistical preferences) of
each residue present in protein. Ideal score S
Ideal
= exp(À0.83 + 1.008 ·
ln(L)); where L is the number of amino acids. S
Ideal
score is compatibility
of the sequence with their 3D structure. It is obtained from total score/
ideal score. S
Ideal
score above 0.45S
Ideal
.
134 M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143
42
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
sequence, 46 and 52% of identity and similarity, respective-
ly) in two conformational states. In the monomer A, the
loop 22 is opened over the FMN binding site and in the
other monomers (B, C, and D), the loop 22 is closed over
FMN binding site. Fig. 6 (A to D) shows the superposition
of Ca atoms of the MtCS with others solved structures.
Fig. 6 demonstrates that MtCS core is similar to other
structures. However, some dislocated regions can be
observed, mainly the H7 in MtCS. In the superposition
of the MtCS with HpCS, it can be observed that the H13
in MtCS is in opposing direction. Furthermore, superposi-
tion of MtCS against other CS structures indicates highest
RMSD for HpCS and ScCS, possibly due to low identity of
these structures with MtCS (31 and 33% of identity, respec-
tively). On the other hand, the AaCS and SpCS structures
present higher identity with MtCS (45 and 46% of identity,
respectively) and therefore a major structural resemblance.
Furthermore, the MtCS structure presents highest resem-
blance with the monomer B, C, and D from SpCS (which
are in the closed conformation—Fig. 6D). The residues
of FMN binding site of complex HpCS:FMN also present
different conformations when compared with MtCS. The
position of residues of active site of HpCS resemble with
the structure of monomer A of SpCS . The FMN binding
site from AaCS and ScCS structures are not defined and
they can not be compared with that region from MtCS.
Fig. 2. MtCS monomer structure. (A) Representation of the monomer MtCS secondary structure elements. (B) MtCS secondary structure representation.
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 135
43
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
This information lead us conclude that in MtCS, the struc -
ture may be in its closed conformation, and in the presence
of FMN, the loop 29 (in MtCS) moves away opening the
EPSP binding site. In the EPSP bound state, the loop 29
from MtCS molecule must close over the ligand. The SpCS
structure has confirmed that the binding of FMN and
EPSP is ordered (Maclean and Ali, 2003), so as Macheroux
et al. (1996, 1998) observed in experiments for EcCS that
oxidized FMN which has a K
D
of about 30 lM, and
decreases to around 20 nM in the presence of EPSP. There-
fore, it is clear that the binding of EPSP blocks any possible
exit of FMN from the active site, and the induced structur-
al changes in the active site allow FMN to make further
interactions with SpCS (Maclean and Ali, 2003).
The residues of FMN and EPSP binding active site were
superposed with the same residues in the two conforma-
tional states of SpCS (Figs. 7A and B). The positions of
many residues are conserved, but there are significant dif-
ferences in the positions of some residues, mainly the
Arg46 and Arg341 at the EPSP binding site and in
Arg110 and Ile317 at FMN binding site. The His11, an
essential residue to CS catalysis (Kitzing et al., 2004) is con-
served in several sequences (Fig. 4) and presents similar
conformation in MtCS and SpCS (in the both SpCS con-
formational states). The His115 in the MtCS structure,
another conserved and important residue to CS catalyzed
reaction (Kitzing et al., 2004), also keeps its position, when
compared with the SpCS open structure. This same resi-
due, when compared to the closed structure of SpCS pre-
sents dislocated approximately 30° outward from the
EPSP binding site (Fig. 7). In SpCS, the His110 (equivalent
the His115 in MtCS) demonstrated therefore some confor-
mational flexibility and interacts with FMN in the open
form but with EPSP in the closed form (Maclean and
Ali, 2003). The closing of the loop 22 on the closed
structure of SpCS can be the cause of this displacement.
The Arg3 41, located in the loop 29 in MtCS (relative to
loop the 22 in SpCS) present the Ca closer to the closed
Fig. 3. MtCS tetrameric structure. The figure represents the MtCS tetrameric structure with its secondary structure content.
Fig. 4. Sequence alignment of several sequences of CS. Alignment was performed with MultAlin program (Corpet, 1988). Numbering is relative to MtCS
primary sequence. Residues shown in red have a correlation of at least 90% with the consensus sequence and are therefore highly conserved. Residues
shown in blue have 50% or more correlation with the consensus sequence and are therefore moderately conserved. Residues in (——) boxes are relative the
active site. Residues in (- - -----)boxes are relative the dimerization interface. Residues in (ÁÁÁÁÁÁÁÁÁ) boxes are relative the tetramerization interface.
Myctub (Mycobacterium tuberculosis); Aquaeo (Aquifex aeolicus); Bacsub (Bacillus subtilis); Laclac (Lactococcus lactis); Strpne (Streptococcus
pneumoniae); Lismon (Listeria monocytogenes); Staaur (Staphylococcus aureus); Brumel (Brucella melitensis); Esccol (Escherichia coli); Saltyp (Salmonella
typhimurium); Yerpes (Yersinia pestis); Haeinf (Haemophilus influenzae); Xilfas (Xylella fastidiosa) Neimen (Neisseria meningitidis); Camjej (Campylobacter
jejuni); Clamur (Chlamydia muridarum).
c
136 M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143
44
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 137
45
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
structure of SpCS, but its side chain is coming from other
direction, occupying the EPSP binding site while in closed
structure of SpCS, it presents dislocated outward from the
site. This effect occurs probably due to the EPSP binding at
its site, therefore in HpCS, this arginine residue is moved
away from the EPSP binding site (data not shown). The dif-
ferent orientations of important residues of active site of
MtCS, as Arg341, Ser342, and Asp343, when compared with
the structure of SpCS, likely occurs due to apo state of MtCS.
The arginine residues 112 and 139, which are involved in
the formation of binding pocket of EPSP by the aliphatic
portions from MtCS structure do not present significant
modifications when compared with SpCS. On the other
hand, the Arg40 of the MtCS is more distended upon the
EPSP binding site than the same arginine in the two SpCS
conformational states, showing that the Arg40 presents
some flexibility. This observation can also be justified by
absence of the EPSP in the MtCS structure. The arginine
residues 46 and 49 that are important to the EPSP binding
(Maclean and Ali, 2003) are localized at the loop 5, which
was not determined with precision probably due to the
absence of interactions of protein with EPSP.
The largest differences in FMN binding site are
observed in the position of Arg110 and Thr319, where
Fig. 5. MtCS dimeric structure and interface. (A) Figure shows the molecular surface of MtCS dimer formed by monomer A (pink) and D (lime). (B)
Representation of the molecular surface on the interface between monomers A (pink) and B (wheat). (C) Representation of the molecular surface on the
interface between monomers A (pink) and D (slate). The figures were generated for Pymol program (DeLano, 2002). (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this paper.)
138 M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143
46
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
in MtCS structure, their lateral chains are shifted out-
ward from the FMN binding site. The side chains of
Ile37 and Lys315 in MtCS are occupying the FMN bind-
ing site. The Ser342 and Asp343 also present different
orientations when compared with SpCS, may be due to
MtCS to be in apo state.
Fig. 6. Superposition of Ca between MtCS and others CS. (A) from A. aeolicus; (B) H. pylori; (C) S. cerevisae, and (D) S. pneumoniae. Structure in light
blue is MtCS, and in light green is monomer B of SpCS. Structure in dark green is monomer A of SpCS; Region in yellow is loop 22 of SpCS of the
monomer B; region in light red is loop 22 of SpCS of the monomer A; Structure in violet is loop 29 of MtCS. H7 and H13 is referent of MtCS. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this paper.)
Fig. 7. Superposition of residues of the (A) FMN binding site and (B) EPSP binding site. Residues in black are from MtCS, yellow are from monomer B
of the SpCS and in orange is from monomer A of the SpCS. (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this paper.)
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 139
47
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.4. Quaternary structure in solution
Gel filtration column was performed to determine the
native molecular weight of MtCS. The results showed that
MtCS is a dimer in solution.
The hydrodynamic properties of the MtCS were estab-
lished by both sedimentation velocity and sedimentation
equilibrium. The samples did not present apparent aggre-
gation and it shows monodispersive behavior. The sedi-
mentation coefficient s normalized to the standard
sedimentation coefficient s
20,w
and extrapolated at 0 mg/
mL (s
0
20;w
Þ is 4.60 ± 0.02 S which is in accordance with a
particle with molec ular mass (MM) of 80–100 kDa, and
of asymmetrical shape. Furthermore, considering the
diffusion coefficient at standard conditions (3.7–3.8 ·
10
À7
cm
2
seg
À1
) determined by dynamic light scattering
and by the application of the Eq. (1), a MM of MtCS
was 80 ± 2 kDa, a value closer of the expected for its dimer
(MM of monomer is 41.7 kDa). The MM derived from the
sedimentation equilibrium was 83 ± 1 kDa and it is also in
accordance with the MM expected for the MtCS dimer.
However, the resultant plot (Fig. 8A, inset) suggests that
s
20,w
increases with the MtCS con centration, instead to
reduce with the viscosity and molecular crowd effect (Laue,
2001). One hypothesis that can explain this behavior of
MtCS is a rapid dimer–tetramer equilibrium ( Rowe,
1977), which should be concentration dependent.
Furthermore, the results of sedimentation equilibrium
also suggest the existence of dimer–tetramer equilibrium
(Fig. 8B). The association constant determined for sedi-
mentation equilibrium is approxim ately 20 mM
À1
, and
therefore, the dissociation constant—K
D
—about 50 M.
Thus, due to very high K
D
, probable the interactions
that form the tetramer are few or/and very weak in
these experimental conditions. SpCS, on the other hand,
also shows dimer–tetramer equilibrium with dissociation
constant of 0.8 lM(Maclean and Ali, 2003). So, we can
conclude that in solution, the MtCS is predominantly
dimer while SpCS is tetramer, despite of their structural
resemblance.
The comparison between the predicted s
20,w
for MtCS
crystal structures and the s
max
for a globular protein rela-
tive MM correspondents (Table 2) shows that the MtCS
in solution can have an asymmetric shape. Through of
the sedimentation equilibrium and sedimentation velocity
data of MtCS, we can considerate the predicted s
0,w
for
future analysis. But, the experimental s
0
20;w
data for MtCS
do not agree with the predicted data for crystallographic
dimer formed for monomers A/D or B/C of this structure.
These ambiguous results suggest that, in solution, the
dimer of MtCS is slightly different from the crystallograph-
ic dimers (A/D or B/C) that form the tetramer, presenting
a more asymmetrical shape.
These differences of the oligomeric states observed for
MtCS when comp ared with other CS structures may not
0
5.0
0.1
5.1
Absorbance (AU)
0
Residuals
50.0-
50.0
06.655.605.654.604.653.6
)mc(suidaR
tM
SC
mpr000,8
mpr000,01
mpr000,21
0
50.0
01.0
51.0
02.0
52.0
03.0
53.0
g(s*)
0.20.010.80.60.40
*s
Mt
CS
Lm/gm6.0
Lm/gm8.0
Lm/gm0.1
1.00.80.60.40.20
5.20
5.00
4.80
4.60
4.40
Mt
CS )Lm/gm(
s
Mt
CS S)20.0±06.4(=
s (S)
Fig. 8. Sedimentation velocity experiments. (A) Sedimentation velocity experiments: The g (s*) distributions were fitted using the Origin (Microcal
Software) with a Gaussian giving apparent sedimentation coefficients s* for MtCS at different protein concentrations (see Section 2 for details). Inset: plots
of s
20,w
versus protein concentration fitted by linear regression to calculate the s
0
20;w
: 4.60 ± 0.04 Svedberg. Despite the data were well fitted by only one
Gaussian suggesting the existence of only one MtCS specie in solution, the positive slop suggests that the MtCS undergoes an oligomerization process. (B)
Sedimentation equilibrium experiments. The figure shows the best fits of experimental data for 450 lg/mL of MtCS at 8000, 10000, and 12000 rpm with
the self-association methods (see Section 2). The random distribution of the residuals (bottom panel) indicates that the fit is satisfactory. The fitness of the
data obtained at other protein concentration presented similar results (data not shown). Sedimentation equilibrium data of MtCS agree with a dimer
structure with 83 ± 1 kDa in equilibrium with a tetrameric form with an affinity constant of 20 ± 5 mM
À1
.
Table 2
Comparison between the predicted s
20,w
for MtCS crystal structure and
the s
max
estimated for a globular protein
Oligomeric state s
20,w
(S) s
max
(S)
Monomer 3.18 ± 0.05 4.20
Dimer 5.42 ± 0.06 6.70
Tetramer 9.02 ± 0.06 10.70
140 M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143
48
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
be justified by number of saline bridges and hydrogen
bonds in the dimer interface area, which form the tetramer.
Table 3 shows the analysis of the tetramerization interface
for five microorganisms. It can be observed that the num-
ber of saline bridges varies is between 0 and 6 and the num-
ber of hydrogen bonds is between 6 and 22 in the analyzed
structures. Analysis of contact area between dimer s reveals
that all structures present contact interfaces above of
4000 A
˚
, but SpCS present a tetramerization interface of
12,132.0 A
˚
2
. Therefore contacts between dimers that form
the tetramer of SpCS are larger than those observed in
the other structures. The most striking difference between
MtCS and other CS structures is the area percentage of
polar atoms, which participate in the interface regions.
MtCS present 53% of hydrophobic area (referent to the
carbon atoms) while the other CS struc tures present a min-
imum 60%. Therefore, we can affirm that the interface
between the dimers that forms tetramer in MtCS is more
polar an d hydrophilic when compared with other CS struc-
tures. How ever, this result can not justify the predomi-
nance of dimer in solution, but it shows that the
hydrophobic effect has a minor participation in the forma-
tion of tetramer in MtCS structure. On the other hand, in
some way, a minor hydrophobic surface can contribute to
predominance of dimer in solut ion. Furthermore, a
hypothesis to explain the low presence of tetramer in solu-
tion is that MtCS tetramerization may involve small initial
contacts, which lead to some conformational changes
increasing the contact area between dimers, resulting on
MtCS tetramerization. The initial contacts can be concen-
tration-dependent and/or still due to the presence of some
ligand, which may trigger the conformational changes
resulting in the oligomerization (Fig. 9).
4. Conclusion
Here, we report the first crystal structure and analysis in
solution of the chorismate synthase from M. tuberculosis.
The CD analyses show that MtCS secondary structures
content agree with the CS structures from other organisms.
Furthermore, MtCS thermal stability experiments show
two well defined transitions, suggesting a complex unfold-
ing mechanism. The monomeric and tetrameric structures
of MtCS are similar to other CS structures. Many of con-
served residues in the primary sequence of the active site of
CS present conserved conformations in all structures
solved, but some residues have large conformational
changes when compared to other structures. All differences
and resemblances observed in the MtCS structure when
compared to the SpCS and HpCS structures may be due
to absence of any ligand in the active site from MtCS.
The determination of MtCS structure in the presence of
FMN, EPSP, and chorismic acid may confirm these evi-
dences. The intricate association between two monomer s
lanoitamrofnoC
segnahC
hgiH
noitartnecnoc
ro
gnidnibdnagiL
Mt
CS
dimer
tM
SC
remarteT
Fig. 9. MtCS oligomerization mechanism. Initial contacts on the MtCS dimer can be induced for protein concentration or ligand binding, would lead to
some conformational changes increasing the dimers contact area and resulting on MtCS tetramerization.
Table 3
Contact analysis of interface area between the dimers that form the CS tetramer
Structure Mycobacterium
tuberculosis
Streptococcus
pneumoniae
Helicobacter
pylori
Aquiflex
aelicus
Saccharomyces
cerevisae
Overall contact surface (A
˚
2
) 5027.0 12132.0 5305.0 4425.0 5745.0
Contact surface apolar (A
˚
2
) (%) 2661.0 (53) 7303.0 (60) 3369.0 (63) 2658.0 (60) 3653.0 (63)
Contact surface hydrophilic polar (A
˚
2
) (%) 2338.0 (47) 4825.0 (40) 1936 (37) 1737.0 (40) 2092.0 (37)
Saline bridges (number) 2 6 1 2 0
Hydrogen bonds (number) 24 40 26 27 22
The contact surface was calculated by AREA/AREAIMOL (CCP4, 1994) and saline bridges and hydrogen bonds were identified for protein–protein
interaction (http://www.biochem.ucl.ac.uk/bsm/PP/server), on the basis of principles described by Jones and Thornton (1996).
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 141
49
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
of the MtCS leads to dimer organization, connecting to
form a tetramer, which may be described as a dimer of
dimers. In solution, the MtCS must be mainly in dimeric
state that is slightly different from crystallographic dimer,
which forms the tetramer. Analysis of all CS led us to con-
clude that all CS must present quaternary structure in solu-
tion in the dimeric or tetr americ forms or in equilibrium of
them.
Despite of the CS structure presented here is in apo
form, the structural data obtained are enough relevant.
This is the first structure in the apo form that presents
almost all residues of active site, excepting the residues
49–52. In this form, it is possible to observe several move-
ments of the important residues of active site of CS. Fur-
thermore, this is the first CS predominantly dimeric in
structure, which can be due to the amount of interaction
between two dimers, which form the tetrameric structure.
The PDB accession code for the apo crystal structure of
Chorismate synthase from Mycobacterium tuberculosis is
1ZTB.
Acknowledgments
We thank Fundac¸a
˜
o de Amparo a
`
Pesquisa do Estado de
Sa
˜
o Paulo (SMOLBNet, Proc. 01/07532-0 and 03/12472-2),
Conselho Naci onal de Desenvolvimento Cientı
´
fico e
Tecnolo
´
gico and Coordenac¸a
˜
o de Aperfeic¸oamento de
Pessoal de
´
vel Superior for fellowships and financial
support. We thank the LNLS technical staff for assistance.
References
Ahn, H.J., Yoon, H.J., Lee, B., Suh, S.W., 2004. Crystal structure of
chorismate synthase: a novel FMN-binding protein fold and function-
al insights. J. Mol. Biol. 336, 903–915.
Azevedo Jr., W.F., Canduri, F., Oliveira, J.S., Basso, L.A., Palma, M.S.,
Pereira, J.H., Santos, D.S., 2002. Molecular model of shikimate kinase
from Mycobacterium tuberculosis. Biochem. Biophys. Res. Commun.
295 (1), 142–148.
Basso, L.A, Silva, L.H.P., Fett-neto, A.G., Azevedo Jr., W.F., Moreira,
I.S., Calixto, J.B., Astolfi Filho, S., Santos, R.R., Soares, M.B.P.,
Santos, D.S., 2005. The use of biodiversity as source of new chemical
entities against defined molecular targets for treatment of malaria,
tuberculosis, and T-cell mediated diseases—A review. Memo
´
rias do
Instituto Oswaldo Cruz 100 (6), 475–506.
Bentley, R., 1990. The shikimate pathway—metabolic tree with many
branches. Crit. Rev. Biochem. Mol. Biol. 25, 307–384.
Bo
¨
hm, G., Muhr, R., Jaenicke, R., 1992. Quantitative analysis of protein
far UV circular dichroism spectra by neural networks. Protein Eng. 5,
191–195.
Bornemann, S., Ramjee, M.K., Balasubramanian, S., Abell, C., Coggins,
J.R., Lowe, D.J., Thorneley, R.N., 1995. Escherichia coli chorismate
synthase catalyzes the conversion of (6S)-6-fluoro-5-enolpyruvylshik-
imate-3-phosphate to 6-fluorochorismate. Implications for the enzyme
mechanism and the antimicrobial action of (6S)-6-fluoroshikimate. J.
Biol. Chem. 270, 22811–22815.
Bornemann, S., Lowe, D.J., Thorneley, R.N., 1996. The transient kinetics
of Escherichia coli chorismate synthase: substrate consumption,
product formation, phosphate dissociation, and characterization of a
flavin intermediate. Biochemistry 35 (30), 9907–9916.
Bornemann, S., Lawson, D.M., Thorneley, R.N., 2003. A branch point in
chorismate synthase research. Structure 11, 1463–1465.
Bru
¨
nger, A.T., 1992. X-PLOR, a System for Crystallography and NMR.
Yale University Press, London.
Collaborative Computational Project No. 4., 1994. The CCP4 suite:
programs for proteins crystallography, Acta Crystallogr., D 50, 760–763.
Corpet, T.F., 1988. Multiple sequence alignment with hierarchical
clustering. Nucleic Acid Res. 16 (22), 10881–10890.
Dosselaere, F., Vanderleyden, J.A., 2001. A metabolic node in action:
chorismate-utilizing enzymes in microorganisms. Crit. Rev. Microbiol.
27, 75–131.
DeLano, W.L., 2002. The PyMOL Molecular Graphics System. CA
DeLano Scientific.
Dias, M.V.B., Ely, F., Canduri, F., Pereira, J.H., Frazzon, J., Basso, L.A.,
Palma, M.S., De Azevedo, W.F., Santos, D.S., 2004. Crystallization
and preliminary X-ray diffraction analysis of chorismate synthase
Mycobacterium tuberculosis. Acta Crystallogr., D. Biol. Crystallogr.
60, 2003–2005.
Duncan, K., 2004. Identification and validation of novel drug targets in
tuberculosis. Curr. Pharm. Des. 10 (26), 3185–3194.
Edelhock, H., 1967. Spectroscopic determination of tryptophan and
tyrosine in protein. Biochemistry 6, 1948–1954.
Fitzpatrick, T.B., Killer, P., Thomas, R.M., Jelesarov, I., Amrhein, N.,
Macheroux, P., 2001. Chorismate synthase from the hyperthermophile
thermotoga maritima combines thermostability and increased rigidity
with catalytic and spectral properties similar to mesophilic counter-
parts. J. Biol. Chem. 276, 18052–18059.
Gandy, M., Zumla, A., 2002. The resurgence of disease: social and
historical perspectives on the ’new’ tuberculosis. Soc. Sci. Med. 55 (3),
385–396.
Garcı
´
a de la Torre, J., Huertas, M.L., Carrasco, B., 2000. Calculation of
hydrodynamic properties of globular proteins from their atomic-level
structure. Biophys. J. 78, 719–730.
Gill, S.C., von Hippel, P.H., 1989. Calculation of protein extinction
coefficients from amino acid sequence data. Anal. Biochem. 182,
319–326.
Goldberg, R.J., 1953. Sedimentation in the ultracentrifuge. J. Phys. Chem.
57, 194–202.
Hill, R.K., Newkome, G.R., 1969. Stereochemistry of chorismic acid
biosynthesis. J. Am. Chem. Soc. 91, 5893–5894.
Johnson, M.L., Correia, J.J., Yphantis, D.A., Halvorson, H.R., 1981.
Analysis of data from the analytical ultracentrifuge by nonlinear least-
squares techniques. Biophys. J. 36, 575–588.
Jones, S., Thornton, J.M., 1996. Principles of protein–protein interactions.
Proc. Natl. Acad. Sci. USA 93, 13–20.
Jones, T.A., Zou, J.Y., Cowan, S.W., 1991. Improved methods for
building protein models in electron density maps and the location of
errors in these models. Acta Crystallogr., A 47, 110–119.
Kitzing, K., Auweter, S., Amrhein, N., Macheroux, P., 2004. Mechanism
of chorismate synthase. Role of the two invariant histidine residues in
the active site. J. Biol. Chem. 279 (10), 9451–9461.
Laskowski, R.A., MacArthur, M.W, Moss, D.S., Thornton, J.M., 1994.
PROCHECK v.3.0—Program to Check the Stereochemistry Quality
of Protein structures—Operating Instructions.
Laue, T.M., 2001. Biophysical studies by ultracentrifugation. Curr. Opin.
Struct. Biol. 11, 579–583.
Leslie, A.G.W., 1992. MOSFLM version 6.11 for processing image plate
and CCD data.
Macheroux, P., Bornemann, S., Ghisla, S., Thorneley, R.N., 1996. Studies
with flavin analogs provide evidence that a protonated reduced FMN
is the substrate-induced transient intermediate in the reaction of
Escherichia coli chorismate synthase. J. Biol. Chem. 271 (42), 25850–
25858.
Macheroux, P., Schonbrunn, E., Svergun, D.I., Volkov, V.V., Koch,
M.H., Bornemann, S., Thorneley, R.N., 1998. Evidence for a major
structural change in Escherichia coli chorismate synthase induced by
flavin and substrate binding. Biochem. J. 335, 319–327.
Macheroux, P., Schmid, J., Amrhein, N., Schaller, A., 1999. A unique
reaction in a common pathway: mechanism and function of choris-
mate synthase in the shikimate pathway. Planta 207, 325–334.
142 M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143
50
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Maclean, J., Ali, S., 2003. The structure of chorismate synthase reveals a
novel flavin binding site fundamental to a unique chemical reaction.
Structure 11, 1499.
McRee, D.E., 1999. XtalView/Xfit—a versatile program for manipulating
atomic coordinates and electron density. J. Struct. Biol. 125, 156–165.
Murshudov, G.N., Vagin, A.A., Dodson, E.J., 1997. Refinement of
macromolecular structures by the maximum-likelihood method. Acta
Crystallogr., D 53, 240–255.
Navaza, J., 2001. Implementation of molecular replacement in AMoRe.
Acta Crystallogr. D. Biol. Crystallogr. 57 (Pt. 10), 1367–1372.
Pereira, J.H., Canduri, F., Oliveira, J.S., Silveira, N.J.F., Basso, L.A,
Palma, M.S., Azevedo Jr., W.F., Santos, D.S., 2003. Structural
bioinformatics study of EPSP synthase from Mycobacterium tubercu-
losis. Biochem. Biophys. Res. Commun. 312 (3), 608–614.
Pereira, J.H., Oliveira, J.S., Canduri, F., Dias, M.V.B., Palma, M.S.,
Basso, L.A., Santos, D.S., Azevedo Jr., W.F., 2004. Structure of
shikimate kinase from Mycobacterium tuberculosis reveals the binding
of shikimic acid. Acta Crystallogr. D. Biol. Crystallogr. 60 (Pt. 12, Pt.
2), 2310–2319.
Polikarpov, I., Perles, L.A., de Oliveira, R.T., Oliva, G., Castellano, E.E.,
Garratt, R.C., Craievich, A., 1998. Set-up and experimental parameters
of the protein crystallography beam line at the Brazilian National
Synchrotron Laboratory. Nucl. Instrum. Methods Phys. Res. A 405,
159–164.
Quevillon-Cheruel, S., Leulliot, N., Meyer, P., Graille, M., Bre-
mang, M., Blondeau, K., Sorel, I., Poupon, A., Janin, J., van
Tilbeurgh, 2004. Crystal structure of the bifunctional chorismate
synthase from Saccharomyces cerevisiae. J. Biol. Chem. 279,
619–625.
Rowe, A.J., 1977. The concentration dependence of transport processes: a
general description applicable to the sedimentation, translational
diffusion, and viscosity coefficients of macromolecular. Biopolymers
16, 2595–2611.
Silveira, N.J.F., Ucho
ˆ
a, H.B., Pereira, J.H., Canduri, F., Basso, L.A.,
Palma, M.S., Santos, D.S., Azevedo Jr., W.F., 2005. Molecular models
of protein targets from Mycobacterium tuberculosis. J. Mol. Model. 11
(2), 160–166.
Stafford, W.F., 1994. Boundary analysis in sedimentation velocity
experiments. Methods Enzymol. 240, 478–501.
Ucho
ˆ
a, H.B, Jorge, G.E., Silveira, N.J.F., Ca
ˆ
mera, J.C., Canduri, F.,
Azevedo Jr., W.F., 2004. Parmodel: a web server for automated
comparative modeling of proteins. Biochem. Biophys. Res. Commun.
325 (4), 1481–1486.
Viola, C.M., Saridakis, V., Christendat, D., 2004. Crystal structure of
chorismate synthase from aquifex aeolicus reveals a novel beta
alpha beta sandwich topology. Proteins: Struct. Funct. Genet. 54,
166–169.
Welch, G.R., Cole, K.W., Gaertner, F.H., 1974. Chorismate synthase of
Neurospora crassa: a flavoprotein. Arch. Biochem. Biophy. 165, 505–
518.
WHO, 2004. World Health Organization. WHO Report, Geneva, Swit-
zerland. WHO/CDS/TB/2004.331.
M.V.B. Dias et al. / Journal of Structural Biology 154 (2006) 130–143 143
51
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.3 Molecular models of tryptophan synthase from Mycobacterium tuberculosis
complexed with inhibitors
Marcio V. B. Dias; Fernanda Canduri; Nelson J. F. da Silveira; Clarissa M. Czekster; Luiz
A. Basso; Mário S. Palma; Diógenes S. Santos; Walter F. de Azevedo Jr.Cell
Biochemistry and Biophysics (ISSN 1085-9195), v. 44, p. 375-384, 2006.
Neste trabalho é apresentada a modelagem molecular da proteína triptofano sintase
de M. tuberculosis em complexo com 6 inibidores análogos do estado de transição desta
enzima.
Para a modelagem molecular da triptofano sintase de M. tuberculosis, foi utilizado
como ponto de partida (templates), as estruturas cristalográficas da triptofano sintase de S.
typhimurium em complexo com os respectivos inibidores. Para realizar a modelagem
molecular foi utilizado o programa MODELLER (SALI; BLUNDELL, 1993). As duas
subunidades da triptofano sintase foram modeladas separadamente e foram gerados 1000
modelos para cada subunidade e para cada complexo. Os modelos finais foram
selecionados baseando-se em análise estereoquímica. A análise da afinidade de ligação
entre o complexo proteína:ligante foi realizada baseando-se na contribuição individual de
afinidade de ligação de cada átomo do ligante, utilizando o programa SCORE (WANG et
al., 1998). Além disso foi levado em consideração o número de ligações de hidrogênio e a
área de contato entre o ligante e a proteína. Os modelos gerados apresentam boa qualidade
estereoquímica e são bastante similares a seus templates. A comparação dos sítios ativos
dos complexos mostra que muitas interações de hidrogênio são conservadas entre os
modelos obtidos e os templates, assim como muitas interações entre os próprios inibidores,
apesar de estes apresentarem algumas interações únicas. Análise da área de contato entre a
52
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
proteína e seus inibidores mostra que a triptofano sintase de M. tuberculosis apresenta uma
maior área de contato do que a de S. typhimurium, entretanto, análise das interações
favoráveis para a ligação, mostra que a triptofano sintase S. typhimurium apresenta uma
afinidade de ligação (score) maior do que a triptofano sintase de M. tuberculosis. Desta
forma, algumas alterações nos átomos dos inibidores estudados podem aumentar a
afinidade destes pela triptofano sintase de M. tuberculosis.
53
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
O
RIGINAL
A
RTICLE
© Copyright 2006 by Humana Press Inc.
All rights of any nature whatsoever reserved.
1085-9195/(Online)1559-0283/06/44:375–384/$30.00
Cell Biochemistry and Biophysics
375 Volume 44, 2006
INTRODUTION
Tryptophan synthase (TRPS) is a target for develop-
ment of vaccines and drugs. The interaction between
Mycobacterium tuberculosis TRPS (MtTRPS) and in-
hibitors derived from the arylthioalkyl-phosphonated
inhibitors (e.g., indole 3-propanol phosphate [IPP]) is
not known. In this work, we show that IPP and their
analogs (or derivatives) present strong interaction with
MtTRPS and that they can inhibit this protein efficiently.
The development of new therapies as vaccines and
drugs is vital in developing country, where a majority of
the population die because of infectious diseases, such
as tuberculosis, AIDS, cholera, and malaria, among oth-
ers (1). Among these diseases, tuberculosis deserves
special attention because it is the main cause of death in
these countries. M. tuberculosis, the main causative
agent of tuberculosis, infects approx 8 million individu-
als and causes 2 million deaths annually (2).
Bacteria, fungi, apicomplexan parasites, and plants syn-
thesize aromatic compounds, and this synthesis is essential
for their survival (3,4). Therefore, the enzymes responsible
for synthesis of aromatic compounds are considered targets
Molecular Models of Tryptophan Synthase
From Mycobacterium tuberculosis Complexed With Inhibitors
Marcio Vinicius Bertacine Dias,
1
Fernanda Canduri,
2
Nelson José Freitas da Silveira,
1
Clarissa Melo Czekster,
3
Luis Augusto Basso,
4
Mário Sérgio Palma,
5
Diógenes Santiago Santos,
6,
* and Walter Filgueira de Azevedo, Jr.
4,
*
1
Programa de Pós-Graduação em Biofísica Molecular-Departamento de Física, UNESP, São José do Rio Preto, Brazil;
2
Departamento de Morfofisiologia—CCBS—UFMS, Campo Grande-MS, 79070-900, Brazil;
3
Rede Brasileira de Pesquisas em
Tuberculose, Grupo de Microbiologia Molecular e Funcional, Centro de Biotecnologia, UFRGS, Porto Alegre-RS, 91501-970,
Brazil;
4
Faculdade de Biociências—PUCRS, Porto Alegre-RS, 90619-900, Brazil;
5
Laboratory of Structural Biology and
Zoochemistry, CEIS/Department of Biology, Institute of Biosciences, UNESP, Rio Claro, SP 13506-900, Brazil; and
6
Centro de Pesquisas em Biologia Molecular e Funcional/PUCRS, Avenida Ipiranga, 6681,
Tecnopuc, Partenon 90619-900, Porto Alegre, RS, Brazil
Abstract
The development of new therapies against infectious diseases is vital in developing countries. Among infec-
tious diseases, tuberculosis is considered the leading cause of death. A target for development of new drugs is
the tryptophan pathway. The last enzyme of this pathway, tryptophan synthase (TRPS), is responsible for con-
version of the indole 3-glycerol phosphate into indol and the condensation of this molecule with serine-produc-
ing tryptophan. The present work describes the molecular models of TRPS from Mycobacterium tuberculosis
(MtTRPS) complexed with six inhibitors, the indole 3-propanol phosphate and five arylthioalkyl-phosphonated
analogs of substrate of the α-subunit. The molecular models of MtTRPS present good stereochemistry, and the
binding of the inhibitors is favorable. Thus, the generated models can be used in the design of more specific
drugs against tuberculosis and other infectious diseases.
Index Entries: Tryptophan synthase; Mycobacterium tuberculosis; molecular modeling; drug design; structural
bioinformatics.
*
Author to whom all correspondence and reprint requests should
be addressed. E-mail: walter[email protected] or diogenes@ pucrs.br
54
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
for development of new therapies, because these com-
pounds are absent in animals. Thus, the inhibition of amino
acid biosynthesis is a demonstrated mode of action for
important classes of inhibitors. Glyphosate and 6-fluoro-
shikimate, for example, inhibit biosynthesis of aromatic
compounds in bacteria, fungi, apicomplexan parasites, and
plants (5,6). Enzymes involved specifically in tryptophan
biosynthesis have been proposed as potential herbicide
and vaccine targets (7,8). The last enzyme of the pathway of
tryptophan biosynthesis is TRPS, and the following rea-
sons led to selection of this enzyme as a potential target. For
one, all organisms that synthesize tryptophan are known to
do so by a single route, such that inhibition of this pathway
should dramatically reduce tryptophan levels. TRPS is also
one of the most studied enzymes of those involved in the
biosynthesis of amino acids (9). It catalyzes two distinct
reactions by separate polypeptide chains, referred to as α-
and β-subunits. The α-subunit catalyzes the cleavage of
indole 3-glycerol phosphate (IGP) to yield an indole and
glyceraldehyde-3-phosphate, whereas the β-subunit cat-
alyzes the condensation of the indole with serine to pro-
duce tryptophan (Fig. 1) (10). The three-dimensional (3D)
structure of the TRPS from Salmonella typhimurium
(StTRPS) with several inhibitors has been deposited in the
376 Dias et al.
Cell Biochemistry and Biophysics Volume 44, 2006
Fig. 1. Reaction scheme of the TRPS. (A) Reaction of the α-subunit. (B) Reaction of the β-subunit.
55
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Protein Data Bank (PDB). Among them are IPP (11) and
five arylthioalkyl-phosphonated transition state analogs
that inhibit the α-subunit (12). IPP is an analog of IGP that
inhibits the α-subunit with a K
i
value of 15 µM, and it was
the starting point for inhibitor design strategy of the
arylthioalkyl-phosphonate (7). The arylthioalkyl-phos-
phonated inhibitors were designed to mimic the transi-
tion state formed during the α-reaction of the enzyme,
and as expected, they have affinities much greater than
that of the natural substrate IGP or its nonhydrolyzable
analog IPP. These inhibitors are ortho-substituted arylth-
ioalkyl-phosphonate derivatives that have an sp
3
-
hybridized sulfur atom, designed to mimic the putative
tetrahedral transition state at the C3 atom of the indole,
and they lack the C2 atom to allow for higher conforma-
tional flexibility (7,12). These phosphonated inhibitors
have been synthesized because of efforts to develop her-
bicidal and antimicrobial agents, and they have been
tested for inhibition of enzyme activity in in vitro assay
and for herbicidal activity in a biological assay (7). These
phosphonated compounds possess enzyme inhibitory
and herbicidal activities with micromolar IC
50
values (7).
The present work describes six molecular models of
MtTRPS, the complex with IPP, and the complexes with
five arylthioalkyl-phosphonated analogs bound in α-sub-
unit and pyridoxal 5- phosphate (PLP) in the β-subunit.
MATERIALS AND METHODS
Molecular Modeling
For modeling of the MtTRPS, we used Parmodel, a Web
server for automated comparative modeling of proteins
(13) based on the program MODELLER (14). The atomic
coordinates of the crystallographic structures of StTRPS
were used as starting models. For the complex with IPP,
we used PDB access code 1QOP (11) and for five com-
plexes with arylthioalkyl-phosphonated derivatives we
used the following PDB access codes: lC29, IC8V, 1CX9,
1C9D, and 1CW2 (12). The arylthioalkyl-phosphonated
inhibitors are 4-(2-hydroxyphenylthio)-1-butenylphos-
phonic acid, 4-(2-hydroxyphenylthio)butylphosphonic
acid, 4-(2-aminophenylthio) butylphosphonic acid, 4-(2-
hydroxy-5-fluorophenylthio)butylphophonic acid and
4-(2-hydroxyphenylsylfinyl)butylphosphonic acid. The
complexes are designated by numbers, with 1 for
TRPS:IPP and 2 to 6 for the five arylthioalkyl-phospho-
nates, respectively. The chemical structures of inhibitors
are shown in Fig. 2. All the cited compounds bind to the α-
subunit. Moreover, the β-subunit was modeled in complex
with PLP for all of the models. However, the models were
generated separately for the α- and β-subunit. The atomic
coordinates of all water were removed from the TRPS tem-
plates. Several slightly different models can be calculated
by varying the initial structure. In total, 1000 models were
generated for each subunit of tryptophan synthase and
each conformational state, and the final models were
selected based on stereochemical quality. Subsequently, we
performed the junction of the two subunits, and the com-
plexes were optimized using MODELLER (14). The opti-
mization of the complexes was carried out by the use of
the variable target function method using methods of con-
Molecular Models of TRPS 377
Cell Biochemistry and Biophysics Volume 44, 2006
Fig. 2. Molecular formulae of TRPS inhibitors. (1) indol-
propanol-phosphate; (2) 4-(2-hydroxyphenylthio)-1-butenyl-
phosphonic acid; (3) 4-(2-hydroxyphenylthio)butylphos-
phonic acid; (4) 4-(2-aminophenylthio)butylphosphonic acid;
(5) 4-(2-hydroxy-5-fluorophenylthio)butylphophonic acid;
and (6) 4-(2-hydroxyphenylsylfinyl)butylphosphonic acid.
56
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
jugate gradients and molecular dynamics with simulated
annealing. All modeling processes were performed on a
Beowulf cluster, with 16 nodes (B16/AMD Athlon 1800+;
BioComp, São José do Rio Preto, SP, Brazil).
Analysis of the Model
The overall stereochemical quality of the final models
for the complexes of MtTRPS were assessed by the pro-
gram PROCHECK (15). Molecular models were super-
posed using the program LSQKAB from CCP4 (15). The
cutoff for hydrogen bonds and salt bridges was 3.5 Å.
The contact surfaces for the binary complexes were cal-
culated using AREAIMOL and RESAREA (15). The root
mean squared deviation (rmsd) differences from ideal
geometries for bond lengths and bond angles were cal-
culated with X-PLOR (16). The G-factor is essentially just
log-odds score based on the observed distributions of
the stereochemical parameters. It was computed for the
following properties: torsion angles (the analyses pro-
vided the observed distributions of ϕ-ψ, χ
1
χ
2
, χ–1, χ–3,
χ–4, and
ω
values for each of the 20 amino acid types)
and covalent geometry (for the main-chain bond lengths
and bond angles). These average values were calculated
using PROCHECK (15). The 3D profile measures the
compatibility of a protein model with its sequence, and
it was calculated using Verify 3D program (17,18).
Analysis of Binding Affinity
of Protein–Ligand Complexes
For an estimative the absolute binding affinity of the
protein–ligand complexes, we used the program SCORE
(19). According to this method, the binding affinity of the
ligand can be decomposed to the contribution of individ-
ual atoms. Each ligand atom obtains a score, called the
atomic binding score, indicating its role in the binding
process. The program reads the structure, assigns atom
types and parameters, performs the calculation, and gives
the dissociation constant of the given protein–ligand com-
plex. The computational results are outputted into a text
file in which the detailed information of each ligand atom,
including the atomic binding score, is tabulated.
RESULTS AND DISCUSSION
The sequence alignment of StTRPS (template) with
MtTRPS (target) is shown in Fig. 3. The sequence of α-
subunit of MtTRPS shows 27% of identity with the
sequence of α-subunit of StTRPS, whereas the sequence
of β-subunit of MtTRPS shows 53% identity with the
sequence of β-subunit of StTRPS.
Quality of the Model
Ramachandran plots for the six StTRPS structures
solved by crystallography were generated to compare
the overall stereochemical quality of six MtTRPS mod-
els. Analysis of the Ramachandran plots indicated that
the models present a minimum of 93.4% of the residues
in the most favored regions and a maximum of 0.2% of
the residues in the disallowed regions, whereas struc-
tures solved by crystallography present a minimum of
91.6% of the residues in the most favored regions and
also a maximum of 0.2% of the residues in the disal-
lowed regions (Table 1). The Verify 3D values, the aver-
age G-factor, and rmsd values of bond lengths and
bond angles are shown in Table 2. From analysis of the
overall stereochemical quality of the molecular models,
we feel that it is appropriate for structural studies.
Overall Description
The structures of six molecular models were super-
posed with their templates. Only Cα was considered in
the superposition. The values of RMSD for the superpo-
sitions can be seen in Table 2. These results show that
MtTRPS models are most similar to the StTRPS structure,
as expected. The models of MtTRPS contain 270 residues
in the α-subunit (residues 1–8 were omitted) and 422 in
the β-subunit (residues 1–23 were omitted). The model of
MtTRPS complexed with IPP is presented in Fig. 4.
The structure of the α-subunit follows the 8-fold α/β
barrel motif first observed in triosephosphate isomerase
(TIM) (20,21). The α/β barrel structure can be described
in terms of a canonical form in which the molecule is
built up from eight repeating supersecondary structural
units each comprised of a β-strand followed by an α-
helix. The eight helices, connected to both ends of the
strands by loops, pack closely in parallel around the
periphery of the barrel.
The β-subunit of the crystal structure of StTRPS and
the molecular models of MtTRPS reveal the presence of
two domains of nearly equal size, denominated N-
domain and C-domain (21). The core of the N-domain is
formed by four strands, with three helices packed on one
side of the sheet, and a fourth helix (helix 6) packed on the
other side. The part of the N-domain constitutes a domain
that has very few interactions with the rest of the protein
and is named in the COMM-domain (21). The C-domain
is comprised of a six-strand β-sheet surrounded by α-
helices. The core of the C-domain contains a central β-
sheet sandwiched between layers of parallel packed
helices, much like the core of the N-terminal domain. The
β-sheet contains five parallel strands and one antiparallel.
The interface between α- and β-subunits is mostly
hydrophobic in character, and many residues of the
interface between two subunits are conserved in almost
all sequences. Approximately 1295 Å
2
of the surface
area of the α-subunit and 1285 Å
2
of the β-subunit are
buried at the α/β interface.
Interaction Between Protein and Its Inhibitors
All inhibitors studied in this work bind to the active
site of the α-subunit. Potential hydrogen bonding inter-
378 Dias et al.
Cell Biochemistry and Biophysics Volume 44, 2006
57
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
actions and their relative distances from active site
residues for all different inhibitors to MtTRPS can be
seen in Fig. 5. The majority of the interactions are com-
mon to all inhibitors; only some are unique. The
inhibitors form several electrostatic interactions with
different parts of the enzyme. The alkylphosphonate
portion of the phosphonated inhibitors and portions of
the phosphate of IPP extend approximately at a right
angle with the phenyl ring, and the oxygen of phospho-
nate and phosphate groups forms hydrogen bonds with
the hydroxyl group of Ser240 and the main-chain nitro-
gens of Gly190, Gly219, Gly239, and Ser240. The other
residue that interacts with inhibitors is Asp68, which
forms a hydrogen bond with hydroxyl group for com-
Molecular Models of TRPS 379
Cell Biochemistry and Biophysics Volume 44, 2006
Fig. 3. Sequence alignment for TRPS from M. tuberculosis and S. typhimurium. (A) sequence alignment for α-subunit. (B)
Sequence alignment for β-subunit. The alignments were performed with the program ClustalW (26), with the meaning of
the special characters in the last line as follows: (*) positions that have a single, fully conserved; (:) one of the following
“strong” groups is fully conserved; and (.) one of the following “weaker” groups is fully conserved.
58
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
plexes 2, 3, 5, and 6, or the amino group for complex 4
bonded to the phenyl ring of the inhibitors, or with the
indole nitrogen atom of complex 1. Asp68 is thought to
play a crucial role in catalysis of the α-reaction and the
intersubunit communication interface (22). Compounds
1, 4, 5, and 6 present a third interaction with main chain
nitrogens of Ser240. In compound 1, the introduction of
the double bond does not disturb the pattern of the
hydrogen bond between the enzyme and inhibitor,
although rigid conformation is imposed. The atom of
the fluorine introduced in the ring of compound 5 does
not interact by polar interaction with any residue of the
protein. On the other hand, the sulfoxide oxygen of
compound 6 forms one hydrogen bond interaction with
the hydroxyl group of Tyr181. By comparison of the
binding site of the six ligands in the models of MtTRPS,
we may conclude that it superimposes significantly well
(figure not shown). The major differences occur in
inhibitor 4 in overall positioning of its alkyl chain that is
associated with an altered orientation of the phosphate
group position relative to the other inhibitors. Owing to
the orientation of the phosphate group in compound 4,
it presents a weaker interaction between the inhibitor
and the backbone amine of Gly190 compared with oth-
ers inhibitors (3.4 Å).
The presence of different ligands in the α-subunit
does not cause any significant change in the binding site
of PLP in the β-subunit, despite that TRPS is an
allosteric protein and binding of the compounds of the
α-subunit is able to cause structural change in the β-
subunit (11).
Analysis of the Interaction Between the Molecular
Models and Inhibitors
Table 3 shows the analysis of the interaction between
the molecular models of MtTRPS and the molecular
structure of StTRPS with the six compounds studied in
the present work. Roughly, all ligands have high affinity
for MtTRPS and StTRPS. They present from six to eight
intermolecular hydrogen bonds, contact area above 179
Å
2
, and positive score value. Therefore, it can be said
that the high affinity between the TRPS and IPP or
arylthioalkyl-phosphonated inhibitors is related to
directional hydrogen bonds and ionic interactions, as
well as shape complementarity of the contact surface of
the partners (23–25).
From analysis (Table 3), compound 5 presents greater
contact area with MtTRPS and seven hydrogen bonds,
but does not present greater amount of favorable bind-
ing (score value). This suggests that compound 5 can be
buried in the protein, but the binding is not completely
favorable. Compound 6 presents eight hydrogen bonds,
but the score value is the lowest. Therefore, this com-
pound must present binding that is not favorable for
interaction with protein. In contrast, compound 2 pre-
sents greater score value and thus a larger amount of
favorable binding with protein and a large contact area
(199 Å). Therefore, this compound presents higher affin-
ity with the protein compared with other compounds.
In biological activity assay, compound 2 presented
greater herbicide activity against Arabidopsis thaliana (7).
Accordingly, based on modeling results obtained and
experimental results, compound 2 can also inhibit
MtTRPS. Inhibition studies of MtTRPS against com-
pound 2 and other inhibitors studied here may confirm
these predictions.
CONCLUSIONS
The molecular models of MtTRPS generated by
MODELLER program possess good stereochemical
quality. Furthermore, they show that the interaction
with phosphonates and IPP is favorable as observed for
StTRPS. Therefore, the molecular models can be used
in the development of more specific drugs against
tuberculosis and other infectious diseases, because the
380 Dias et al.
Cell Biochemistry and Biophysics Volume 44, 2006
Table 1
Analysis of Ramachandram Plots
Most Additional Generously Disallowed
Model favored (%) allowed (%) allowed (%) (%)
1 96.8 (94.6) 5.5 (5.1) 0.5 (0.2) 0.2 (0.2)
2 93.2 (94.7) 6.2 (5.1) 0.5 (0.2) 0.0 (0.0)
3 93.8 (92.7) 5.7 (7.1) 0.4 (0.2) 0.2 (0.0)
4 94.0 (91.6) 5.5 (8.2) 0.4 (0.0) 0.2 (0.2)
5 93.4 (91.6) 6.0 (8.0) 0.5 (0.4) 0.0 (0.0)
6 93.4 (93.6) 6.0 (6.2) 0.5 (0.2) 0.0 (0.0)
Values in parentheses are for templates.
59
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Molecular Models of TRPS 381
Cell Biochemistry and Biophysics Volume 44, 2006
Table 2
Analysis of Quality of Models of MtTRPS and Templates
rmsd from
3D profile
a
G-factor
b
ideal geometry Superposition
Torsion Covalent Bond Bond rmsd
Complex Total score Ideal score S
ideal
score angles geometry Global lengths (Å) angles (°) (Å)
1 301.80 (281.35) 303.60 (298.97) 0.99 (0.94) –0.24 (0.12) –0.25 (–0.21) –0.24 (017) 0.021 (0.008) 2.519 (0.02) 0.128
2 298.95 (338.03) 303.60 (298.97) 0.98 (1.13) –0.27 (0.03) –0.30 (0.43) –0.28 (0.20) 0.021 (0.010) 2.515 (1.70) 0.232
3 302.55 (330.26) 303.60 (298.97) 1.00 (1.10) –0.27 (0.05) –0.28 (0.40) –0.28 (0.20) 0.021 (0.010) 2.686 (1.80) 0.256
4 293.33 (324.04) 303.60 (298.97) 0.97 (1.08) –0.28 (0.02) –0.29 (0.45) –0.28 (0.20) 0.021 (0.009) 2.686 (1.71) 0.248
5 298.52 (323.71) 303.60 (298.97) 0.98 (1.08) –0.30 (0.05) –0.33 (0.41) –0.30 (0.20) 0.022 (0.010) 2.745 (1.83) 0.297
6 297.38 (327.72) 303.60 (298.97) 0.98 (1.10) –0.14 (0.01) –0.44 (0.42) –0.25 (0.18) 0.022 (0.010) 2.785 (1.77) 0.249
Values obtained for the templates used for modeling are presented in parentheses.
a
Total score is the sum of the 3D-one-dimensional scores (statistical prefer-
ences) of each residue present in protein. Ideal score S
Ideal
= exp(–0.83 + 1.008 × ln(L)), where L is the number of amino acids. S
Ideal
score is compatibility of the
sequence with the 3D structure. It is obtained from total score/ideal score. S
Ideal
score above 0.45S
Ideal
.
b
Ideally, scores should be above –0.5. Values below –1.0 may need investigation.
60
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
382 Dias et al.
Fig. 4. Ribbon diagram of the α- and β-subunits of MtTRPS complexed with IPP and PLP generated by Molmol (27).
Fig. 5. Hydrogen bond pattern between MtTRPS and inhibitors. (1) indol-propanol-phosphate; (2) 4-(2-hydroxyphenylthio)-
1-butenylphosphonic acid; (3) 4-(2-hydroxyphenylthio)butylphosphonic acid; (4) 4-(2-aminophenylthio)butylphosphonic acid;
(5) 4-(2-hydroxy-5-fluorophenylthio)butylphophonic acid; and (6) 4-(2-hydroxyphenylsylfinyl)butylphosphonic acid.
61
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
examination of the binding site for arylthiophos-
phanate inhibitor presents interactions that could be
exploited to improve affinity with the protein. The
replacement of hydrogen atoms along the alkyl chain
with larger atoms or functional group can increase the
affinity or the inhibition of the ligands.
ACKNOWLEDGMENTS
This work was supported by grants from FAPESP
(SMOLBNet 01/07532-0, 02/04383-7, 03/12472-2,
04/00217-0), CNPq, CAPES, and Instituto do Milênio
(CNPq-MCT) PRONEX, FAPERGS. W.F.A., M.S.P.,
L.A.B., and D.S.S. are researchers for the Brazilian
Council for Scientific and Technological Development
from Brazil (CNPq).
REFERENCES
1. Trouiller, P., Torreele, E., Olliaro, P., et al. (2001) Drugs for
neglected diseases: a failure of the market and a public
health failure? Trop. Med. Int. Health 6, 945–951.
2. World Health Organization. Global Tuberculosis Control.
WHO Report 2001. Geneva, Switzerland, WHO/CDS/
TB/2001.287
3. Dosselaere, F. and Vanderleyden, J. (2001) A metabolic
node in action: chorismate-utilizing enzymes in microor-
ganisms. Crit. Rev. Microbiol. 27, 75–131.
4. Roberts, F., Roberts, C. W., Johnson, J. J., et al. (1998)
Evidence for the shikimate pathway in apicomplexan par-
asites. Nature 393, 801–805.
5. Schonbrunn, E., Eschenburg, S., Shuttleworth, W. A., et al..
(2001) Interaction of the herbicide glyphosate with its tar-
get enzyme 5-enolpyruvylshikimate 3-phosphate synthase
in atomic detail. Proc. Natl. Acad. Sci. U.S.A. 98, 1376–1380.
6. Daves, G. M, Barrett-Bee, K. J. Jude, D. A., et al. (1994) (6S)-
6-Fluoroshikimic acid, an antibacterial agent acting on the
aromatic biosynthetic pathway. Antimicrob. Agents
Chemother. 38, 403–406.
7. Finn, J., Langevine, C., Birk, I., Nicherson, K., and
Rodaway, S. (1999) Rational herbicide design by inhibition
of tryptophan biosynthesis. Bioorg. Med. Chem. Lett. 9,
2297–2302.
8. Smith, D. A., Parish, T., Stoker, N. G., and Bancroft, G. J.
(2001) Characterization of auxotrophic mutants of
Mycobacterium tuberculosis and their potential as vaccine
candidates. Infect. Immun. 69, 1142–1150.
9. Hyde, C. C. and Miles, E. W. (1990) The tryptophan syn-
thase multienzyme complex: exploring structure-function
relationships with X-ray crystallography and mutagene-
sis. Biotechnology 8, 27–32.
10. Pan, P., Woehl, E., and Dunn, M. F. (1997) Protein architec-
ture, dynamics and allostery in trytophan synthase chan-
neling. Trends Biochem. Sci. 22, 22–27.
11. Weyand, M. and Schlichting, I. (1999) Crystal structure of
wild-type tryptophan synthase complexed with the nat-
ural substrate indole-3-glycerol phosphate. Biochemistry
38, 16,469–16,480.
12. Sachpatzidis, A., Dealwis, C., Lubetsky, J. B., Liang, P. H.,
Anderson, K. S., and Lolis, E. (1999) Crystallographic
studies of phosphonate-based alpha-reaction transition-
state analogues complexed to tryptophan synthase.
Biochemistry 38, 12,665–12,674.
13. Uchoa, H. B., Jorge, G. E., da Silveira, N. J. F., Camera, J. C.,
Jr., Canduri, F., and de Azevedo, W. F., Jr. (2004) Parmodel:
a web server for automated comparative modeling of pro-
teins. Biochem. Biophys. Res. Commun. 325, 1481–1486.
14. Sali, A. and Blundell, T. L. (1993) Comparative protein
modelling by satisfaction of spatial restraints. J. Mol. Biol.
234, 779–815.
15. Collaborative Computational Project No. 4 (1994) The
CCP4 suite: program for protein crystallography. Acta
Crystallogr. D 50, 760–763.
16. Brünger, A. T. (1992) X-PLOR version 3.1: a system for
crystallography and NMR. Yale University Press, New
Haven, CT.
17. Bowie, J. U., Luthy, R., and Eisenberg, D. (1991) A method
to identify protein sequences that fold into a known three-
dimensional structure. Science 253, 164–170.
18. Luthy, R., Bowie, J., and Eisenberg D. (1992) Assessment of
protein models with three-dimensional profiles. Nature
356, 83–85.
19. Wang, R., Liu, L., Lai, L., and Tang, Y. (1998) SCORE: a
new empirical method for estimating the binding affinity
of a protein-ligand complex. J. Mol. Model. 4, 379–394.
20. Hyde, C. C., Ahmed, S. A., Padlan, E. A., Miles, E. W., and
Davies, D. R. (1988) Three-dimensional structure of the
tryptophan synthase α
2
β
2
multienzyme complex from
Salmonella typhimurium. J. Biol. Chem. 263, 17,857–17,871.
21. Banner, D. W., Bloomer, A. C., Petsko, G. A., et al. (1975)
Structure of chicken muscle triose phosphate isomerase
determined crystallographically at 2.5 Å resolution using
amino acid sequence data. Nature 255, 609–614.
22. Rhee, S., Miles, E. W., Mozzarelli, A., and Davies, D. R.
(1998) Cryocrystallography and microspectrophotometry
of a mutant (alpha D60N) tryptophan synthase alpha 2
beta 2 complex reveals allosteric roles of alpha Asp60.
Biochemistry 37, 10,653–10,659.
Molecular Models of TRPS 383
Cell Biochemistry and Biophysics Volume 44, 2006
Table 3
Interaction of Between Protein and Inhibitors
a
Number
hydrogen Contact
Model bonds area (Å
2
) Score
1 7 (7) 200.0 (167.0) 7.39 (7.76)
2 6 (6) 199.0 (176.0) 7.98 (8.29)
3 6 (6) 197.0 (191.0) 7.70 (7.94)
4 7 (7) 192.0 (172.0) 6.61 (6.95)
5 7 (7) 211.0 (185.0) 6.51 (7.60)
6 8 (7) 179.0 (175.0) 5.52 (7.04)
a
Values in parenthesis are for templates.
62
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
23. De Azevedo, W. F., Jr., Mueller-Dieckmann, H. J., Schulze-
Gahmen, U., Worland, P. J., Sausville, E., and Kim, S. H.
(1996) Structural basis for specificity and potency of a
flavonoid inhibitor of human CDK2, a cell cycle kinase.
Proc. Natl. Acad. Sci. U.S.A. 93, 2735–2740.
24. De Azevedo, W. F., Jr., Canduri, F., Dos Santos, D. M., et al.
(2003) Structural basis for inhibition of human PNP by
immucillin-H. Biochem. Biophys. Res. Commun. 309,
922–927.
25. Pereira, J. H., Canduri, F., de Oliveira, J. S., et al. (2003)
Structural bioinformatics study of EPSP synthase from
Mycobacterium tuberculosis. Biochem. Biophys. Res. Commun.
312, 608–614.
26. Thompson, J. D., Higgins, D. G., and Gibson, T. J. (1994)
CLUSTAL W: improving the sensitivity of progressive
multiple sequence alignment through sequence weight-
ing, positions-specific gap penalties and weight matrix
choice. Nucleic Acids Res. 22, 4673–4680.
27. Koradi, R., Billeter, M., and Wüthrich, K. (1996) MOL-
MOL: a program for display and analysis of macromolec-
ular structures. J. Mol. Graph. 14, 51–55.
384 Dias et al.
Cell Biochemistry and Biophysics Volume 44, 2006
63
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.4 Effects of the magnesium and chloride ions and shikimate on the structure of
shikimate kinase from Mycobacterium tuberculosis.
Marcio V. B. Dias; Lívia M. Faím; Igor B. Vasconcelos; Jaim S. de Oliveira; Luiz A.
Basso; Diógenes S. Santos; Walter F. de Azevedo Jr. Acta Crystallographica Section F
Structural Biology and Crystallization Communications (ISSN 1744-3091), 63, p. 1-6,
2007.
Neste trabalho é apresentado o efeito dos íons magnésio e cloro e do ácido
chiquímico sobre a estrutura da chiquimato quinase de M. tuberculosis (peso molecular de
18583 Da).
A Chiquimato quinase de M. tuberculosis foi clonada, superexpressada e purificada
pelo grupo de Pesquisas do Prof. Dr. Diógenes S. Santos da PUC – Porto Alegre – RS.
Para a obtenção de cristais da chiquimato quinase de M. tuberculosis em complexo com
ácido chiquímico e ADP, na ausência de Mg
2+
, a proteína foi dialisada contra tampão Tris-
HCl, 50 mM, pH 8,0 e concentrada a 14 mg.mL
-1
e utilizada inicialmente a condição de
cristalização estabelecida por Dhaliwal et al. (2004). Os melhores cristais foram obtidos
em uma condição composta for Tris-HCl, 0,1 M, pH 8, 17% de PEG 1500 e de cloreto de
lítio, 0,5-0,7 M. Por outro lado, a condição de cristalização na qual foram obtidos os
melhores cristais para o complexo da chiquimato quinase de M. tuberculosis em complexo
com ADP e magnésio era composta por Tris-HCl, 0,1 M, 20% PEG 3350 e 0,1 M de
MgCl
2
.6H
2
O e a concentração da proteína era de 17 mg.mL
-1
. A coleta de dados de
difração de raios X foi realizada no LNLS. O cristal obtido para a chiquimato quinase de
M. tuberculosis em complexo com ADP e ácido chiquímico é similar aos descritos na
literatura e é pertencente ao grupo espacial P3
2
21. Este cristal difratou a 1,93 Å de
resolução e apresenta um monômero na unidade assimétrica. Os cristais da chiquimato
64
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
quinase de M. tuberculosis em complexo com Mg
2+
e ADP são pertencentes ao grupo
espacial P2
1
2
1
2
1
e difrataram a 2,8 Å de resolução, e apresentam na unidade assimétrica
quatro monômeros, que formam um tetrâmero com simetria 222. As estruturas para os
conjuntos de dados dos cristais obtidos foram determinadas por substituição molecular
utilizando como modelo de busca a estrutura da chiquimato quinase de M. tuberculosis em
complexo com MgADP e ácido chiquímico (PEREIRA et al., 2004). O refinamento
cristalográfico das estruturas foi realizado pelo programa REFMAC 5 (MURSHUDOV;
VAGIN; DODSON, 1997) e a inspeção visual e adição de moléculas de água foi efetuada
pelo programa XtalView/Xfit (McREE et al., 1999). A estrutura da chiquimato quinase de
M. tuberculosis em complexo com ADP e ácido chiquímico, na ausência do íon magnésio,
foi comparada com estrutura da chiquimato quinase em complexo com MgADP e ácido
chiquímico determinada por Pereira et al., 2004, que é a única estrutura que foi observado
o íon magnésio quando o ácido chiquímico está presente. Pela comparação destas duas
estruturas pôde-se observar o efeito que o íon magnésio causa sobre a estrutura da
chiquimato quinase e sobre o ácido chiquímico. Nesta estrutura foram observadas
alterações na posição da cadeia lateral de importantes resíduos do sítio ativo da enzima e
alterações nos grupos hidroxilas da molécula de ácido chiquímico. Com relação à estrutura
da chiquimato quinase de M. tuberculosis em complexo com magnésio e ADP pôde-se
observar as possíveis interferências do ácido chiquímico sobre a conformação da
chiquimato quinase, uma vez que esta molécula não foi adicionada na condição de
cristalização. Além disso, em um dos monômeros do tetrâmero, a molécula de cloro, que
representa um importante papel na catálise da chiquimato quinase, não foi observada
(CERASOLI et al., 2003). Assim, este monômero apresentou alterações estruturais não
observadas nos outros monômeros. Estas alterações conformacionais foram descritas e
acreditamos que elas podem ser causadas pela ausência do íon cloreto na estrutura da
chiquimato quinase de M. tuberculosis.
65
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
protein structure communications
Acta Cryst. (2007). F63, 1–6 doi:10.1107/S1744309106046823 1
Acta Crystallographica Section F
Structural Biology
and Crystallization
Communications
ISSN 1744-3091
Effects of the magnesium and chloride ions and
shikimate on the structure of shikimate kinase from
Mycobacterium tuberculosis
Marcio Vinicius Bertacine Dias,
a
´
via Maria Faı
´
m,
a
Igor Bordin
Vasconcelos,
b
Jaim Simo
˜
es de
Oliveira,
b
Luiz Augusto Basso,
b
Dio
´
genes Santiago Santos
b
* and
Walter Filgueira de Azevedo Jr
c
*
a
Programa de Po
´
s-Graduac¸a
˜
o em Biofı
´
sica
Molecular, Departamento de
´
sica, UNESP,
Sa
˜
o Jose
´
do Rio Preto, SP 15054-000, Brazil,
b
Pontifı
´
cia Universidade Cato
´
lica do Rio Grande
do Sul, Centro de Pesquisa em Biologia
Molecular e Funcional, Porto Alegre, RS, Brazil,
and
c
Faculdade de Biocie
ˆ
ncias, Pontifı
´
cia
Universidade Cato
´
lica do Rio Grande do Sul,
Av. Ipiranga, 6681 Porto Alegre-RS,
CEP 90619-900, Brazil
Correspondence e-mail: [email protected],
Received 27 September 2006
Accepted 6 November 2006
PDB References: shikimate kinase–ADP–
shikimate complex, 2dfn, r2dfnsf; shikimate
kinase–MgADP complex, 2dft, r2dftsf.
Bacteria, fungi and plants can convert carbohydrate and phosphoenolpyruvate
into chorismate, which is the precursor of various aro matic compounds. The
seven enzymes of the shikimate pathway are responsible for this conversion.
Shikimate kinase (SK) is the fifth enzyme in this pathway and converts
shikimate to shikimate-3-phosphate. In this work, the conformational changes
that occur on binding of shikimate, magn esium and chloride ions to SK from
Mycobacterium tuberculosis (MtSK) are described. It was observed that both
ions and shik imate influence the conformation of residues of the active site of
MtSK. Magnesium influence s the conformation of the shikimate hydroxyl
groups and the position of the side chains of some of the residues of the active
site. Chloride seems to influence the affinity of ADP and its position in the active
site and the opening length of the LID domain. Shikimate binding causes a
closing of the LID domain and also seems to influence the crystallographic
packing of SK. The results shown here could be useful for understanding the
catalytic mechanism of SK and the role of ions in the activity of this protein.
1. Introduction
Mycobacterium tuberculosis, the aetiological agent of tuberculosis
(TB), infects one-third of the world’s population. It is estimated that
1.7 million deaths resulted from TB in 2004, 95% of which occurred in
developing countries (World Health Organization, 2006). The emer-
gence of TB as a public health threat, the high susceptibility of HIV/
TB co-infected patients and the proliferation of multi-drug-resistant
strains have created a need for newer and better drugs for the
treatment of TB.
The shikimate pathway is an attractive target for the development
of herbicides (Coggins, 1998) and antibiotic agents (Davies et al.,
1994) because it is essential in algae, higher plants, bacteria, fungi and
apicomplexan parasites but is absent from mammals (Bentley, 1990;
Roberts et al., 1999). The shikimate pathway is a seven-step biosyn-
thetic route that links the metabolism of carbohydrates to the
synthesis of aromatic amino acids. The shikimate pathway leads to the
biosynthesis of chorismate, which is a precursor of aromatic amino
acids and many other aromatic compounds (Ratledge, 1982). Shiki-
mate kinase (SK; EC 2.7.1.71), the fifth enzyme of this pathway,
catalyzes phosphate transfer from ATP to the carbon-3-hydroxyl
group of shikimate, forming shikimate 3-phosphate (S3P).
SK belongs to the nucleoside monophosphate (NMP) kinase
structural family. SK is a / protein consisting of a central sheet of
five parallel -strands flanked by -helices, with overall topology
similar to that of adenylate kinase (Pereira et al., 2004; Krell et al.,
1998, 2001). A characteristic feature of the NMP kinases is that they
undergo large conformational changes during catalysis (Vonrhein et
al., 1995). The NMP kinases are composed of three domains: the
CORE, which contains a highly conserved phosphate-binding loop
(P-loop), the LID domain, which undergoes substantial conforma-
tional changes upon substrate binding, and the NMP-binding domain,
# 2007 International Union of Crystallography
All rights reserved
electronic reprint
66
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
which is responsible for the recognition and binding of a specific
substrate (Gu et al., 2002).
MgADP induces concerted hinged movements of the shikimate-
binding (SB) and LID domains such that the two domains move
towards each other and towards the active centre of the enzyme in the
presence of this ligand (Gu et al., 2002).
The precise mode of binding of shikimate to MtSK and some
conformational changes upon shikimate binding to MtSK have
recently been proposed (Pereira et al., 2004; Dhaliwal et al., 2004).
The binding of shikimate to MtSK causes a concerted conformational
protein structure communications
2 Dias et al.
Shikimate kinase Acta Cryst. (2007). F63, 1–6
Figure 1
(a) Structure of MtSK in complex with ADP, shikimate and Cl
À
.(b) Tetrameric structure of MtSK in complex with ADP, Mg
2+
and Cl
À
. The monomers A, B, C and D are
represented in green, blue, pink and yellow, respectively. The dark blue and yellow spheres represent the magnesium and chloride ions, respectively. The ADP molecules are
represented as sticks. (c) Representation of the hydrogen-bonding interactions that occur between the ribose hydroxyl groups of ADP molecules bound to monomers A and
B. The distances are shown in A
˚
. Figures were generated with the program MolMol (Koradi et al., 1996).
change of the LID and SB domains towards each other and results in
an additional closure of the active site.
Chloride ions have been shown to weaken the interaction between
shikimate and SK from Erwinia chrysanthemi and to strengthen the
affinity of the enzyme for ADP and ATP (Cerasoli et al., 2003). Thus,
a chloride ion seems to occupy a site crucial for the binding of the
nucleotide substrate in the correct orientation for catalysis.
In this work, we report two crystallographic structures: the MtSK–
ADP–shikimate and MtSK–MgADP complexes. The MtSK–ADP–
shikimate complex has been solved at 1.93 A
˚
resolution. The crystal
structure of MtSK in complex with ADP and Mg
2+
has been solved at
2.8 A
˚
resolution. The data presented here provide an evaluation of
the effects of both Mg
2+
and Cl
À
ions and shikimate on the structure
of MtSK. Thus, these results provide further insight into the roles that
different ions and substrate play in the mechanism of action of MtSK.
2. Materials and methods
2.1. Crystallization
The cloning, expression and purification of MtSK have been
reported elsewhere (Oliveira et al., 2001). Initially, MtSK was
concentrated to 14 mg ml
À1
and dialyzed against 50 mM Tris–HCl
buffer pH 8.0. For the crystallization of MtSK in complex with ADP
and shikimate in the absence of Mg
2+
ion, variations of the crystal-
lization condition established by Dhaliwal et al. (2004) were used.
Crystals were obtained by the hanging-drop vapour-diffusion
method. The well solution contained 0.1 M Tris–HCl buffer pH 8.0,
17% PEG 1500 and 0.5–0.7 M LiCl. The drops had a protein solu-
tion:well solution volume ratio of 1.6:1.0. For the crystallization of the
protein in complex with ADP and Mg
2+
in the absence of shikimate, a
crystallization solution composed of 0.1 M Tris–HCl buffer pH 8.0,
20% PEG 3350 and 0.1 M MgCl
2
.6H
2
O was used. The concentration
of the protein was 17 mg ml
À1
and the drops again had a protein
solution:well solution volume ratio of 1.6:1.0.
electronic reprint
67
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
2.2. Data collection and processing
All data sets were collected at a wavelength of 1.427 A
˚
using a
synchrotron-radiation source (Station PCr, LNLS, Campinas, Brazil;
Polikarpov et al., 1998) and a CCD detector (MAR CCD). Data
collection was performed under cryogenic conditions at a tempera-
ture of 100 K in a cold nitrogen stream generated and maintained
with an Oxford Cryosystem. Prior to flash-cooling, glycerol was
added to the crystallization drop to 20%(v/v). The data sets were
processed using the program MOSFLM (Leslie, 1992) and scaled
with SCALA (Collaborative Computational Project, Number 4,
1994).
2.3. Structure determination
The crystal structures of both complexes were determined by
standard molecular-replacement methods using the program AMoRe
(Navaza, 2001). For both complexes, we used as a search model the
structure of MtSK–MgADP–shikimate (PDB code 1we2; Pereira et
al., 2004). Refinement of the structures was performed using
REFMAC5 implemented in the CCP 4 package (Murshudov et al.,
1997; Collaborative Computational Project, Number 4, 1994). Xtal-
View/Xfit (McRee, 1999) was used for visual inspection and addition
of water molecules. The stereochemical correctness of the models was
checked using PROCHECK (Laskowski et al., 1993). The final atomic
models were superposed using LSQKAB from the CCP4 package
(Collaborative Computational Project, Number 4, 1994). PAR-
MODEL (Ucho
ˆ
a et al., 2004) was used in the final analysis of the
model.
3. Results and discussion
MtSK crystallized in two different space groups depending on the
complex. The crystals of the MtSK–ADP–shikimate ternary complex
were trigonal, space group P3
2
21, and diffracted to 1.93 A
˚
resolution.
The asymmetric unit contains one molecule and the final values of R
and R
free
were 20.2 and 27.0%, respectively. In contrast, the crystals of
the MtSK–ADP–Mg
2+
ternary complex were orthorhombic, space
group P2
1
2
1
2
1
, and diffracted to 2.8 A
˚
resolution. The asymmetric
unit contains four MtSK monomers that form a tetramer and the final
values of R and R
free
were 18.3 and 28.0%, respectively. Table 1
details the data processing, refinement statistics and quality analysis
of the two complexes.
The structures present good geometry, although some residues of
the LID domain are located in disallowed regions of the Rama-
chandran plot. The N-terminal methionine residue is not observed
since it was removed during the MtSK expression in E. coli (Oliveira
et al., 2001). The ten C-terminal residues are disordered in both
structures and have not been included in the final structure.
The folding of the MtSK complexes presented here is similar to
that of those reported previously (Pereira et al., 2004; Krell et al.,
1998; Gu et al., 2002; Dhaliwal et al., 2004). SK is a / protein and
consists of five central parallel -strands flanked by -helices. Fig. 1(a)
shows a ribbon representation of the secondary-structure elements of
MtSK–ADP–shikimate at 1.93 A
˚
resolution.
protein structure communications
Acta Cryst. (2007). F63, 1–6 Dias et al.
Shikimate kinase 3
Figure 3
Superposition of the chains of the structure of the MtSK–MgADP binary complex obtained in the absence of shikimate. (a) Chains A (yellow) and C (light blue), (b) chains B
(green) and C (light blue) and (c) chains B (green) and D (red).
Figure 2
Superposition of the structures of the MtSK–MgADP–shikimate and MtSK–ADP–
shikimate ternary complexes. The C
trace of the MtSK–MgADP–shikimate
complex is presented in green and that of the MtSK–ADP–shikimate trace is
presented in blue. The C atoms of MtSK–MgADP–shikimate and MtSK–ADP–
shikimate are coloured white and yellow, respectively. The Mg
2+
shown in yellow
and the chloride ion shown in turquoise refer to the MtSK–MgADP–shikimate
structure, while the chloride ion in dark blue refers to the MtSK–ADP–shikimate
structure. The figure was generated with the program MolMol (Koradi et al., 1996).
electronic reprint
68
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.1. Structure of the MtSK–ADP–Mg
2+
ternary complex
The asymmetric unit of the MtSK–ADP–Mg
2+
ternary complex
structure contains four monomers forming a homotetramer with 222
symmetry. However, each subunit presents a different conformation
state. In the A and C subunits the LID domain is disordered and
residues 115–123 and 114–123 were thus not included in the final
model.
The structure of the MtSK tetramer is shown in Fig. 1(b). Each
monomer is in contact with the other three, creating an intricate
packing arrangement. The ADP molecules appear to play an
important role in the stabilization of the tetramer, since the hydroxyl
groups of the ribose moiety of ADP of one monomer form hydrogen
bonds with those of a neighbouring monomer (Fig. 1c).
3.2. Influence of Mg
2+
on the structure of MtSK
The absence of Mg
2+
ions seems to have a significant effect on the
position of shikimate and some of the active-site residues, mainly
Asp32 and Asp34 (these are conserved residues in all SKs and are
involved in the binding of Mg
2+
; Fig. 2). In addition, the chloride ion
and ADP molecule also undergo changes in position. Fig. 2 shows a
protein structure communications
4 Dias et al.
Shikimate kinase Acta Cryst. (2007). F63, 1–6
superposition of the MtSK–MgADP–shikimate (Pereira et al., 2004)
and MtSK–ADP–shikimate complexes, showing the structural
changes caused by the absence of the Mg
2+
ion.
In the structure of the MtSK–ADP–shikimate ternary complex, the
3-, 4- and 5-hydroxyl groups of shikimate undergo a shift in their
positions, leading to small differences in the hydrogen-bonding
pattern between shikimate and MtSK. Although the Asp34 residue
interacts with shikimate in both structures, a hydrogen bond between
the OD2 atom of Asp34 and the O3 atom of the shikimate is not
observed in the MtSK structure without Mg
2+
(not shown).
The position of the chloride ion is closer to shikimate in the MtSK–
ADP–shikimate ternary complex than in the MtSK–MgADP–
shikimate complex (Pereira et al., 2004). In the structure of the
MtSK–ADP–shikimate ternary complex, the chloride ion is 2.4 A
˚
away from the O1 atom of the shikimate 3-hydroxyl group, while in
the structure of MtSK–MgADP–shikimate the distance between
these two atoms is 3.4 A
˚
. The absence of the Mg
2+
ion can also cause
alterations in the side chains of the hydrophobic residues Phe49 and
Phe57 located in the SB domain and of Pro118, Ala46 and Ile45
located in the LID domain (Fig. 2).
As established previously for other NMP kinases, the Mg
2+
ion is
expected to play an important role in the transfer of the -phosphate
group of ATP to the nucleophilic 3-OH group of shikimate by an
associative reaction mechanism (Schlichting & Reinstein, 1997;
Bellinzoni et al., 2006). Thus, the differences observed in the MtSK–
ADP–shikimate ternary complex in the absence of Mg
2+
may
contribute to understanding of the MtSK chemical mechanism.
3.3. Influence of chloride ion on the structure of MtSK
In our structure of the MtSK–MgADP binary complex, monomer
C of the MtSK tetramer does not contain chloride ion. This ion seems
to influence the position of the ADP molecule and the position of the
shikimate molecule. This monomer presents a conformation that is
more open than the other monomers (Fig. 3). Furthermore, the LID
domain of monomer C has a mean B factor that is higher than those
of the other monomers (not shown). The ADP molecule bound to
monomer C is shifted by approximately 3.0 A
˚
in the active site (Fig. 3)
compared with the other monomers of the MtSK tetramer or with
other structures of MtSK in complex with ADP. These results are in
Figure 4
Pattern of water bonding in the active site of MtSK. (a) Structure at 1.93 A
˚
resolution (chloride ion and shikimate present, magnesium ion absent), (b) monomer B of the
tetramer (chloride and magnesium ions present, shikimate absent), (c) monomer C of the tetramer (magnesium ion present, shikimate and chloride ion absent). The figures
were generated using PyMOL (DeLano, 2004).
electronic reprint
69
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
agreement with the kinetic and spectroscopic data previously
obtained by Cerasoli et al. (2003). This work shows that the chloride
ion increases the stability of the E. chrysanthemi SK structure and
that this same ion also influences the affinity of ADP for SK. Thus,
the increase in the protein stability may be a consequence of chloride
favouring the SK structure in its closed state.
The chloride ion bound to the MtSK active site seems to be part of
an intricate network formed of water molecules, residues of the LID
and SB domains, ADP and shikimate. This network of interactions
appears to cause closure of the structure (Fig. 4). In all structures of
MtSK reported so far and in the three monomers of the MtSK
tetramer, an interaction network involving three water molecules is
formed in the active site (Fig. 4). These water molecules bridge the
interactions between ADP, chloride and magnesium ion and MtSK.
In our MtSK–ADP–shikimate ternary complex, water 306 occupies a
similar position to the Mg
2+
ion. In the absence of chloride ion, this
phenomenon does not occur and furthermore the formation of this
intricate network that can induce the opening of the structure is
avoided.
3.4. Influence of shikimate in the structure of MtSK
The crystals of the MtSK–MgADP binary complex in the absence
of shikimate were obtained in space group P2
1
2
1
2
1
, which has not
been previously described for MtSK. The asymmetric unit presents
four MtSK monomers that form a tetramer. The absence of shikimate
from the crystallization conditions influences the crystal packing and
also the conformation of the SB and LID domains of MtSK, as
observed by Pereira et al. (2004) and Dhaliwal et al. (2004). In
accordance with this, Gan and coworkers recently solved the apo-
MtSK and MtSK–shikimate binary complex structures and suggested
that shikimate binding defines the conformational change of the
protein that arises when shikimate is bound: the LID domain is
ordered and closes over the bound shikimate (Gan et al., 2006).
4. Conclusion
Here, we report two structures of shikimate kinase from M. tuber-
culosis: the MtSK–ADP–shikimate and MtSK–MgADP complexes.
In the former, we observe the effect of the Mg
2+
ion on the structure
and in the latter we observe the effect of shikimate on the crystal
packing and on the structure of MtSK. The Mg
2+
ion seems to
influence the position of the hydroxyl groups of the shikimate
molecule and some of the residues of the active site of MtSK. The
crystal structure of the MtSK–MgADP complex was solved in space
group P 2
1
2
1
2
1
, which has not previously been described for MtSK. In
this space group, the MtSK presents a tetramer with 222 symmetry, in
which the ribose moiety of the ADP molecule seems to play an
important role in the stabilization of the tetramer and the contacts
between the monomers, which occur mainly in the LID-domain
region. However, one monomer of the tetramer does not contain
chloride ion. The absence of this ion seems to cause large changes in
the position of the ADP molecule and also causes a large opening of
MtSK. This information is accordance with the results obtained by
Cerasoli et al. (2003), which shows the importance of the chloride ion
in the stability and the alignment of the substrates in the active site of
SK.
We hope that the results described here will shed light on the
structural changes of MtSK upon binding of substrate(s) that will be
useful for the understanding of the catalytic mechanism and for
structure-based drug design of novel inhibitors that may be potential
anti-mycobacterial agents.
This work was supported by grants from FAPESP (SMOLBNet,
Proc. 01/07532-0, 03/12472-2, 04/00217-0), CNPq, CAPES and Insti-
tuto do Mile
ˆ
ncio (CNPq-MCT), DSS, WFA (CNPq, 300851/98-7) and
LAB (CNPq, 520182/99-5) are researchers of the Brazilian Council
for Scientific and Technological Development.
References
Bellinzoni, M., Haouz, A., Grana, M., Munier-Lehmann, H., Shepard, W. &
Alzari, P. M. (2006). Proteins, 15, 1–5.
Bentley, R. (1990). Crit. Rev. Biochem. Mol. Biol. 25, 307–384.
Cerasoli, E., Kelly, S. M., Coggins, J. R., Lapthorn, A. J., Clarke, D. T. & Price,
N. C. (2003). Biochim. Biophys. Acta, 1648, 43–54.
Coggins, J. R. (1998). In Herbicides and Plant Metabolism, edited by A. Dodge.
Cambridge University Press.
protein structure communications
Acta Cryst. (2007). F63, 1–6 Dias et al.
Shikimate kinase 5
Table 1
Crystallographic data, refinement statistics and analysis of the quality of MtSK
structures.
Values in parentheses are for the outermost shell.
MtSK–ADP–shikimate MtSK–MgADP
Crystallographic data
Unit-cell parameters
a (A
˚
) 63.3 60.2
b (A
˚
) 63.3 62.2
c (A
˚
) 91.6 170.6
Space group P3
2
21 P2
1
2
1
2
1
No. of measurements 76792 97059
No. of independent reflections 16017 17057
Completeness (%) 96.9 (91.5) 99.4 (96.9)
R
sym
(%) 8.9 (58.8) 12.1 (58.0)
Redundancy 4.8 5.7
Refinement statistics
Resolution range (A
˚
) 35.16–1.93 57.17–2.80
Reflections used for refinement 15130 15670
Final R factor‡ (
¨
%) 20.2 18.3
Final R
free
§ (%) 27.0 28.0
Correlation coefficient (%) 95.2 94.3
B values (A
˚
2
)
Main chain 31 33
Side chain 34 36
ADP 21 27
Shikimate 32
Waters 39 30
Quality of structure
Three-dimensional profile} S = 88.07, IS = 74.95,
S/IS = 1.18IS
S = 343.59, IS = 294.34,
S/IS = 1.17IS
Ramachandran plot
Favoured 95.6 84.8
Additionally allowed 2.9 13.3
Generously allowed 0.7 0.8
Disallowed 0.7 1.1
R
sym
= 100
P
IðhÞÀhIðhÞij=
P
IðhÞ, where I (h) is the observed intensity and hI(h)i is
the mean intensity of reflection h over all measur ements of I(h). R factor =
100
P
jF
obs
À F
calc
j=
P
F
obs
, the sums being taken over all reflections with F/(F)>
2(F). § R
free
is the R factor for 10% of the data that were not included during
crystallographic refinement. } The ideal score measures the compatibility of a protein
model with its sequence, using a 3D profile. Each residue position in the 3D model is
characterized by its environment and is represented by a row of 20 numbers in the profile.
These numbers are the statistical preferences (called 3D-1D scores) of each of the 20
amino acids for this environment. Environments of residues are defined by three
parameters: the area of the residue that is buried; the fraction of side chain area that is
covered by polar atoms (O and N) and the local secondary structure. The 3D profile score
S for the compatibility of the sequence with the model is the sum, over all residue
positions, of the 3D-1D scores for the amino-acid sequence of the protein. For 3D protein
models known to be correct, the 3D profile score S for the amino-acid sequence of the
model is high, by contrast, the profile score S for the compatibility of a wrong 3D protein
model with its sequence is often low. When this method is used to verify a structure, the
raw compatibility score alone is difficult to interpret. In this case it is necessary to
compare the score to those obtained using structures known to be correct, we use the
Ideal Score (IS), that is calculated from the length of the protein. The IS is determined by
IS ¼ exp½À0:83 þ 1:008  lnðLÞ. Where L is the length of the sequence. Severely
misfolded structures typically have scores less than 0.45 IS. A score near or above IS
indicates a reliable structure.
electronic reprint
70
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Collaborative Computational Project, Number 4 (1994). Acta Cryst. D50,
760–763.
Davies, G. M., Barret-Bee, K. J., Jude, D. A., Lehan, M., Nichols, W. W. &
Pinder, P. E. (1994). Agents Chemother. 38, 403–406.
DeLano, W. L. (2004). The PyMOL Molecular Graphics System. DeLano
Scientific, San Carlos, CA, USA.
Dhaliwal, B., Nichols, C. E., Ren, J., Lockyer, M., Charles, I., Hawhins, A. R. &
Stammers, D. K. (2004). FEBS Lett. 574, 49–54.
Gan, J., Gu, Y., Li, Y., Yan, H. & Ji, X. (2006). Biochemistry, 45, 8539–8545.
Gu, Y., Reshetnikova, L., Li, Y., Wu, Y., Yan, H., Singh, S. & Ji, S. (2002). J.
Mol. Biol. 319, 779–789.
Koradi, R., Billeter, M. & Wu
¨
thrich, K. (1996). J. Mol. Graph. 14, 51–55.
Krell, T., Coggins, J. R. & Lapthorn, A. J. (1998). J. Mol. Biol. 278, 983–997.
Krell, T., Maclean, J., Boam, D. J., Cooper, A., Resmini, M., Brocklehurst, K.,
Kelly, S. M., Price, N. C., Lapthorn, A. J. & Coggins, J. (2001). Protein Sci. 10,
1137–1149.
Laskowski, R. A., MacArthur, M., Moss, D. S. & Thornton, J. M. (1993). J.
Appl. Cryst. 26, 283–291.
Leslie, A. G. W. (1992). Jnt CCP4/ESF–EACBM Newsl. Protein Crystallogr. 26.
McRee, D. E. (1999). J. Struct. Biol. 125, 156–165.
Murshudov, G. N., Vagin, A. A. & Dodson, E. J. (1997). Acta Cryst. D53,
240–255.
Navaza, J. (2001). Acta Cryst. D57, 1367–1372.
Oliveira, J. S., Pinto, C. A., Basso, L. A. & Santos, D. S. (2001). Protein Expr.
Purif. 22, 430–435.
Pereira, J. H., Oliveira, J. S., Canduri, F., Dias, M. V. B., Palma, M. S., Basso,
L. A., Santos, D. S. & Azevedo, W. F. Jr (2004). Acta Cryst. D60, 2310–
2319.
Polikarpov, I., Perles, L. A., de Oliveira, R. T., Oliva, G., Castellano, E. E.,
Garratt, R. C. & Craievich, A. (1998). Nucl. Instrum. Methods Phys. Res. A,
405, 159–164.
Ratledge, C. (1982). The Biology of the Mycobacteria, Vol. 1, edited by C.
Ratledge & J. L. Stanford, pp. 185–271. London: Academic Press.
Roberts, F., Roberts, C. W., Johnson, J. J., Kyle, D. E., Krell, T., Coggins, G. H.,
Milhous, W. K., Tzipoki, S., Ferguson, D. J., Chakrabarti, D. & McLeod, R.
(1999). Nature (London), 397, 219–220.
Schlichting, I. & Reinstein, J. (1997). Biochemistry,
36, 9290–9296.
Ucho
ˆ
a, H. B., Jorge, G. E., Silveira, N. J. F., Ca
ˆ
mera, J. C., Canduri, F. &
Azevedo, W. F. Jr (2004). Biochem. Biophys. Res. Commun. 325, 1481–
1486.
Vonrhein, C., Schlauderer, G. J. & Schulz, G. F. (1995). Structure, 3, 483–
490.
World Health Organization (2006). Tuberculosis Fact Sheet No. 104. http://
www.who.int/mediacentre/factsheets.
protein structure communications
6 Dias et al.
Shikimate kinase Acta Cryst. (2007). F63, 1–6
electronic reprint
71
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3.5 Crystallographic studies on the binding of isonicotinyl-NAD adduct to wild-
type and isoniazid resistant 2-trans- Enoyl-ACP (CoA) Reductase from
Mycobacterium tuberculosis
Marcio V. B. Dias; Adriane M. X. Prado; Igor B. Vasconcelos; Valmir Fadel; Luis A.
Basso; Diógenes S. Santos; Walter F. de Azevedo Jr. Artigo aceito no Journal
Structural Biology, 2007.
Neste trabalho são apresentadas as estruturas cristalográficas da enzima Enoil
(ACP) redutase (InhA) (peso molecular de 28528 Da) em complexo com INH-NAD na
forma selvagem e para dois mutantes encontrados em isolados clínicos resistentes a
isoniazida (S94A e I21V). É apresentada também a estrutura do mutante S94A na sua
forma nativa.
As proteínas selvagens e os mutantes S94A e I21V foram expressadas e
purificadas pelo grupo do Prof. Dr. Diógenes S. Santos da PUC – Porto Alegre –RS. Os
cristais da InhA (selvagem e mutantes) em complexo com INH-NAD foram produzidos
seguindo as condições estabelecidas por Dessen et al., 1995 e Rozwarski et al., 1998.
Para a repetição destas condições as proteínas foram dialisadas contra HEPES, 50 mM,
pH 7,5. A condição de cristalização na qual foram obtidos os melhores cristais era
composta por HEPES, 50 mM, pH 7.2, citrato de sódio, 50 mM e 5-10% de MPD. Para
a cristalização do mutante S94A na forma nativa, a proteína foi dialisada contra HEPES,
50 mM, pH 7,5, 10% de glicerol e KCl, 300 mM. A condição na qual foram obtidos os
melhores cristais apresentava citrato de sódio, 100 mM, pH 5,6, acetato de amônio, 200
mM e 20-30% de PEG 4000. Os dados de difração de raios X foram coletados no
LNLS. Os cristais obtidos para a InhA em complexo com INH-NAD são pertencentes
72
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
ao grupo espacial P6
2
22 e os cristais obtidos para a proteína na sua forma nativa são
pertencentes ao grupo espacial P1. Todas as estruturas foram determinadas por
substituição molecular e refinadas pelo programa REFMAC 5.2 (MURSHUDOV;
VAGIN; DODSON, 1997).
Os dados obtidos para o complexo da InhA selvagem com INH-NAD apresenta
uma melhor resolução (2.2 Å) do que os dados anteriormente descritos na literatura (2.8
Å) (ROZWARSKI et al., 1998) . Desta maneira, é possível realizar uma melhor análise
do modo de interação desta droga com o sítio ativo da enzima, e evidenciar alterações
ainda não descritas. Além disso, esta estrutura, devido a sua melhor resolução, é mais
confiável para estudos a posteriori no desenvolvimento de novas drogas contra
tuberculose. As estruturas determinadas para os mutantes S94A e I21V contribuem para
a compreensão do mecanismo de resistência destes mutantes à isoniazida, além de
observar a influência que estas mutações causam sobre o processo de ligação do aducto
INH-NAD no sítio ativo da enzima. A estrutura do mutante S94A na sua forma nativa, é
a primeira estrutura apresentada neste estado, e tem sido obtida em uma nova condição
de cristalização e em um novo grupo espacial (P1). A estrutura apresenta um tetrâmero
na unidade assimétrica, que pode corresponder ao tetrâmero encontrado em solução.
Nesta estrutura, foi possível observar o estado da enzima antes da ligação de qualquer
ligante e, desta forma, pôde dar evidências sobre os movimentos de loops e resíduos que
estão envolvidos no processo de ligação de substratos acil ácidos graxos, da molécula de
NADH ou do aducto INH-NAD.
73
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
1
Crystallographic studies on the binding of isonicotinyl-NAD adduct to wild-type
and isoniazid resistant 2-trans- Enoyl-ACP (CoA) Reductase from Mycobacterium
tuberculosis
Marcio Vinicius Bertacine Dias
a
, Igor Bordin Vasconcelos
b
, Adriane Michele Xavier
Prado
a
, Valmir Fadel
a
, Luiz Augusto Basso
b
, Walter Filgueira de Azevedo Jr.
c*
.
Diógenes Santiago Santos
b*
,
a
Programa de Pós-Graduação em Biofísica Molecular Departamento de Física,
UNESP, São José do Rio Preto, SP 15054-000, Brasil;
b
Centro de Pesquisa em Biologia Molecular e Funcional, Instituto de Pesquisas
Biomédicas, Pontifícia Universidade Católica do Rio Grande do Sul, Porto Alegre, RS,
Brasil;
c
Faculdade de Biociências - Pontifícia Universidade Calica do Rio Grande do Sul,
Av. Ipiranga, 6681. Porto Alegre-RS CEP 90619-900 Brasil.
*
Corresponding authors: diogenes@pucrs.br (D.S. Santos) and
walter.junior@pucrs.br
(W.F. de Azevedo Jr.).
telephone: +55 51 3320-3500 4529
fax: +55 17 51 3220-3629
* Manuscript
74
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
2
Abstract
The resumption of tuberculosis led to an increased need to understand the
molecular mechanisms of drug action and drug resistance, which should provide
significant insight into the development of newer compounds. Isoniazid (INH), the most
prescribed drug to treat TB, inhibits an NADH-dependent enoyl-acyl carrier protein
reductase (InhA) that provides precursors of mycolic acids, which are components of
the mycobacterial cell wall. InhA is the major target of the mode of action of isoniazid.
INH is a pro-drug that needs activation to form the inhibitory INH-NAD adduct.
Missense mutations in the inhA structural gene have been identified in clinical isolates
of M. tuberculosis resistant to INH. To understand the mechanism of resistance to INH,
we have solved the structure of two InhA mutants (I21V and S94A), identified in INH-
resistant clinical isolates, and compare them to INH-sensitive WT InhA structure in
complex with the INH-NAD adduct. We also solved the structure of unliganded INH-
resistant S94A protein, which is the first report on apo form of InhA. The salient
features of these structures are discussed and should provide structural information to
should improve our understanding of the mechanism of action of, and resistance to, INH
in M. tuberculosis. The unliganded structure of InhA allows identification of
conformational changes upon ligand binding and should help structure-based drug
design of more potent antimycobacterial agents.
Keywords: M. tuberculosis; InhA; Crystal structure; Isoniazid; drug resistance
75
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
3
Introduction
Tuberculosis (TB), which is caused mainly by Mycobacterium tuberculosis, is a
global human health emergency that remains the leading cause of mortality among
infectious diseases. It has been estimated that 8.2 million new TB cases occurred
worldwide in the year 2000, with approximately 1.8 million deaths in the same year,
which translates into more than 200 deaths per hour, and more than 95 % of these were
in developing countries (Corbett et al., 2003). In the same year, 3.2 % of the world’s
new cases of TB were multidrug-resistant tuberculosis (MDR-TB), defined as strains
resistant to at least isoniazid and rifampicin (Espinal, 2003; Ormerod, 2005). Treatment
of MDR-TB strains requires the administration of second-line drugs that are more toxic
and less effective, and are given for at least three times as long as, and 100 times as
expensive as basic chemotherapeutic regimens (Pablos-Mendes et al., 2002). More
recently, a survey of the frequency and distribution of extensively drug-resistant (XDR)
TB cases, which are defined as cases in persons with TB whose isolates were resistant
to isoniazid and rifampicin and at least three of the six main classes of second-line
drugs, showed that during 2000-2004, of 17,690 TB isolates, 20% were MDR and 10%
of these were XDR (CDC, 2006). XDR-TB has a wide geographic distribution, poses a
public health threat, is an impediment to TB control, and opens up the possibility that
epidemics of virtually untreatable TB may develop (No authors listed, 2006). New
antimycobacterial agents are thus needed to improve the treatment of MDR- and XDR-
TB, as well as to provide more effective treatment of drug-sensitive TB infection. An
understanding of drug resistance mechanisms in this pathogen should contribute to the
rational design of new chemotherapeutic agents to treat TB.
The modern, standard “short-course” therapy for tuberculosis is based on a four-
drug regimen of isoniazid, rifampicin, pyrazinamide, and ethambutol or streptomycin
76
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
4
for two months, followed by treatment with a combination of isoniazid and rifampicin
for four months (Mitchison, 1985). Isoniazid (INH, isonicotinic acid hydrazide) was
first reported to be effective in the treatment of TB in 1952 (Bernstein et al., 1952) and,
soon after, the first INH-resistant M. tuberculosis strains were isolated (Middlebrook,
1953). Genetic and biochemical studies have shown that the inhA-encoded protein is the
primary target for isoniazid (Banerjee et al., 1994; Quémard et al., 1995; Larsen et al.,
2002; Kremer et al., 2003). InhA was identified as an NADH-dependent 2-trans enoyl-
ACP (acyl carrier protein) reductase enzyme that exhibits specificity for long-chain
thioester substrates. InhA is a member of the mycobacterial type II fatty acid synthase
system (FAS-II), which elongates acyl fatty acid precursors yielding the long carbon
chain of the meromycolate branch of mycolic acids, the hallmark of mycobacteria
(Schroeder
et al., 2002). INH is a pro-drug that is activated by the mycobacterial
catalase-peroxidase enzyme KatG in the presence of manganese ions, NAD(H) and
oxygen (Johnsson and Schultz, 1994; Johnsson et al., 1995; Basso et al., 1996; Zabinski
and Blanchard, 1997). The KatG-produced acylpyridine fragment of isoniazid is
covalently attached to the C4 position of NADH forming an INH-NAD adduct, which,
in turn, forms an inhibitory binary complex with the wild-type (WT) enoyl reductase of
M. tuberculosis (Rozwarski et al., 1998) with an equilibrium dissociation constant
value lower than 0.4 nM (Lei et al., 2000). The isonicotinyl-NAD adduct has been
shown to be a slow, tight-binding competitive inhibitor of WT InhA with an overall
inhibition constant value of 0.75 nM (Rawat et al., 2003).
The mechanism of action of isoniazid is complex, as mutations in at least five
different genes (katG, inhA, ahpC, kasA, and ndh) have been found to correlate with
isoniazid resistance (Schroeder
et al., 2002; Blanchard, 1996; Basso and Blanchard,
1998; Glickman and Jacobs, 2001; Basso and Santos, 2005, Oliveira et al., 2007).
77
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
5
Consistent with InhA as the primary target of INH mode of action, INH-resistant
clinical isolates of M. tub erculosis harboring inhA-structural gene missense mu tations,
but lacking mutations in the inhA promoter region, katG gene and oxyR-ahpC region,
were shown to have higher dissociation constant (K
d
) values for NADH than INH-
sensitive WT InhA, whereas there were only modest differences in the steady-state
parameters (Blanchard, 1996). We have recently reported the crystal structures of binary
complexes formed between NADH and INH-sensitive WT InhA, and INH-resistant
S94A, I47T, and I21V InhA mutant enzymes (Oliveira et al., 2006). Even more
recently, both specialized linkage transduction has been used to introduce S94A single
poin t mu tation within the inhA structural gene and X-ray crystallographic on INH-
resistant S94A InhA protein has been reported (Vilchèze et al., 2006). However, even
though there are several crystal structures of InhA in complex with a variety of ligands,
there has been no report on unliganded InhA structure and, thus, no high resolution
information on the InhA structure before ligand binding. In our efforts to understand the
molecular basis for the reduced inhibition of the INH-NAD adduct to InhA mutants,
here we report co-crystallization of INH-resistant I21V and S94A InhA mutant
enzymes, which were identified in INH-resistant clinical isolates of M. tuberculosis
(Blanchard, 1996; Morlock et al., 2003), with the INH-NAD adduct, and compare them
to the INH-sensitive WT InhA structure. This is the first report on the crystal structure
of the complex formed between INH-resistant I21V InhA and INH-NAD adduct refined
to 2.2 Å. We also report the crystal structure of INH-resistant S94A InhA and INH-
sensitive WT InhA both in complex with INH-NAD adduct to 2.0 and 2.2 Å of
resolu tion, respectively. Moreover, we report, for th e first time, the crystal structur e of
apo INH-resistant S94A InhA refined to 2.15 Å, which shows the protein
conformational changes upon ligand binding. It is hoped that the data presented here
78
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
6
will provide structural insight into an understanding of the drug resistance mechanism,
which, in turn, should aid the rational design of chemical compounds to efficiently
inhibit both INH-resistant and –sensitive InhA enzymes with potential
antimycobacterial activity.
Materials and methods
Crysta llizat ion
WT, I21V and S94A InhA enzymes were expressed and purified to homogeneity
as described elsewhere (Quèrmard et al., 1995; Basso et al., 1998). INH-NAD synthesis
was carried out as described elsewhere (Rozwarski et al., 1998) . Crystals of binar y
complex InhA:INH-NAD were obtained by the hanging-drop vapor–diffusion method
under similar conditions as described by Dessen et al., (1995) and Rozwarski et al.,
(1998). InhA enzymes were dialyzed against 50 mM Hepes, pH 7.5 and concentrated to
5-10mg/mL. Enzyme-inhibitor complexes were obtained at room temperature by
incubating 5-10 mg/mL InhA proteins with NADH, MnCl
2
and isoniazid for 1 hour
using molar ratios of 1:50, 1:10, and 1:100, respectively. The complexes InhA:INH-
NAD were crystallized in hanging droplets containing 1 µL of inhibited-protein solution
and 1 µL of crystallization solution containing 50 mM Hepes, pH 7.2; Sodium citrate
buffer and 5-10% MPD.
The apoenzyme S94A InhA mutant was dialyzed against 50 mM Hepes, pH 7.5,
10% glycerol and 300 mM of KCl. The protein was concentrated to 10 mg/mL and
crystallized in hanging droplets con taining 1 µL of protein solution and 0.5 µL of
crystallization solution containing 100 mM sodium citrate, pH 5.6, 200 mM ammonium
acetate and 20-30% of PEG 4000.
79
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
7
Data collection and processing
All data sets were collected at a wavelength of 1.427 Å using Synchrotron
Radiation Source (Station PCr, LNLS, Campinas, Brazil; and a CCD detector
(MARCCD) (Polikarpov et al., 1998). Data collection was performed in cryogenic
conditions at 100 K in a cold nitrogen stream generated and maintained with an Oxford
Cryosystem. Prior to flash-cooling, glycerol was added to the crystallization drop up to
20 % (v/v). The data sets were processed using the program MOSFLM (CCP4, 1994)
and scaled with SCALA (CCP4, 1994).
Structure determination
The crystal structures were determined by standard molecular replacement
methods using the program AMoRe (Navaza, 1994). Initially, atomic coordinates for
binary complex WTInhA:INH-NAD to 2. resolution were used as search model for
the structure of WT InhA:INH-NAD (PDB access code: 1ZID) (Rozwarski et al., 1998).
Our atomic coordinates for WT InhA:INH-NAD at 2.2Å resolution were then used as
search model to solve the other structures. The structure of apoenzyme S94A InhA
mutant was solved using as search model the structure of binary complex InhA:Genz-
10850 at 2. resolution (PDB access code: 1P44) (Kuo et al., 2003). The atomic
position s, which gen erated the higher correlation coefficien t magnitude obtained from
molecular replacement method, were used for the crystallographic refinement. The
refinements of structures were performed using REFMAC 5.2 program (Murshudov
et
al., 1997). The XtalView/Xfit (McRee, 1999) was used for visual inspection and
addition of water molecules. The water molecules were also checked based on B factor
values. The stereochemistry correctness of the models was checked using PROCHECK
80
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
8
program (Laskowski et al., 1993) and PARMODEL (Uchoa et al., 2004). The final
atomic models were superposed using the program LSQKAB program from CCP4i
package (CCP4, 1994).
Results and discussion
The crystals of apo S94A InhA are triclinic
.
These crystals diffracted to 2.15 Å
resolution. The asymmetric unit presents four monomers forming the characteristic
tetramer of InhA. This structure presents final R-factor and R-free values of 16.2% and
25.5%, respectively. The crystals of WT, I21V and S94A InhA in complex with the
INH-NAD adduct are hexagonal and crystallized in the space group P6
2
22, having one
molecule in the asymmetric unit. Table 1 summarizes the data processing,
crystallographic refinement statistics and structural quality for the four structures
presented here. Analysis of the crystallographic refinement for the structures here
described indicates that WtInhA:NADH:INH, I21V InhaA:NADH-INH and
S94Inha:NADH-INH present difference between R-factor and R-free ranging from 3.9
to 5.5 %, which indicates fairly good overall refinement statistics, and the structure
S94A InhA presents a difference of 9.3 %, below 10% , which makes this structure also
acceptable for structural comparisons. These structures present good geometry, although
some residues are located in regions not permitted in Ramachandran plots.
InhA belongs to the short chain dehydrogenase/reductase (SDR) family of
enzymes. Th e main ch aracteristic of th is family is a polyp eptide b ackbone topology in
wh ich each subunit consists of a single domain with a central core that contains a
Rossmann fold supporting an NADH binding site. The structure displays a α/β folding
consisting of a central β-sheet composed of parallel strands and flanked by α-helices
81
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
9
(Figure 1A). The structure presented here is in accordance with the homotetrameric
quaternary structure in solution determined by analytical size-exclusion
chromatography, and possesses an internal 222 symmetry (figure 1B).
Influence of INH on the structure of MtInhA
The activated form of INH consists of an isonicotinic-acyl group attached
through its carbonyl group to the C4 of the nicotinamide ring, replacing the 4S hydrogen
of NADH, which is the same position involved in the hydride transfer that occurs during
reduction of enoyl-ACP substrates (Quémard et al., 1995).
We compared the structure of INH-sen sitive WT InhA presented here and the
structure previously solved by Roswarski et al., (1998) at 2.7 Å of resolution. Although
Vilchèze et al., (2006) have solved two structures of InhA in complex with NAD-INH
(WT InhA and S94A mutant at 2.0 Å and 1.9 Å at resolution respectively) it was only
observed one structure deposited in the Protein Data Bank (S94A InhA mutant). This
way, the S94A InhA solved by Vilchèze et al., (2006) is argued in related to effect of
this mutation on the binding process of the INH-NAD at a later stage.
The structure of the INH-sensitive WT InhA presented here shows differences
when compared with the structure determined by Roswarski et al., (1998), mainly in
important residues located in the binding site of INH-NAD (Tyr158, Phe149, Trp222,
Leu218 and Phe41) and in the INH-NAD molecule. There is a difference of
approximately 15º in the position of Tyr158 side chain of WT InhA in comparison with
the structure solved by Roswarski et al., (1998), and the Tyr158 side chain moves closer
to INH. Interestingly, a conformational change involving rotation of Tyr158 side chain
upon binding of the enoyl substrate to InhA has been invoked to account for the
observed inverse solvent isotope effect (Parikh et al., 1999). These authors have
82
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
10
proposed that Tyr158 functions as an electrophilic catalyst, stabilizing the transition
state for hydride transfer by hydrogen bonding to the substrate carbonyl. The Phe41 side
chain of WT InhA:INH-NAD structure presented here underwent 30º torsion as
compared to the structure solved by Roswarski et al., (1998). This residue appears to be
important in anchoring the adenine moiety of NADH. The side chain of Phe149 and
Leu218 residues move closer to INH molecule (Figure 2A). Although it could be argued
that these differences are due to the lower resolution of the previously published
structure, these residues appear to play an importance role in the enzyme-ligand
interaction and should be considered in structure-based drug design.
The main effect of INH moiety of INH-NAD adduct on WT InhA is a 90º
rotation of Phe149 side chain (Figure 2B), which provides room to accommodate the
INH moiety. Owing to this rotation, a water molecule, which has been identified in the
InhA structures without the isonicotinic-acyl group (Oliveira et al., 2006), was removed
from the InhA active site (not shown in figure 2). Furthermore, it is observed alterations
in the Met155, Leu218 and Ile215, Met199, probably due to alteration in the side chain
position of Phe149. The Ile215 moves away from the active site due to rotation of the
side chain of the Tyr158 (Figure 2B). The side chain of Met199 underwent a rotation of
approximately 3 away from the active site. Two water molecules are present in the
structure of InhA:INH-NAD complex, but absent from the structures between WT and
S94A, I21V and I47T mutants InhA proteins in complex with NADH (Oliveira et al.,
2006). The isonicotinic group also causes the expulsion of one water molecule next to
C4 of NADH molecule. This causes a rotation of approximately 20º of nicotinamide
group of NADH molecule, when compared with other InhA:NADH structures, but the
phosphates, ribose and adenine are highly superposed among the analyzed structures.
83
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
11
This change results in a more snugly fit of the INH moiety of INH-NAD in the active
site of InhA.
There is one intermolecular hydrogen bond that is conserved in the present
structures and in the structure reported by Roswarski et al., (1998) involving oxygen
atom from Gly14 and O2 from the pentose of NADH, which strongly indicates the
importance of this hydrogen bond for the interaction of NADH and the InhA.
Influence of mutations I21V and S94A on INH-NAD binding to InhAs
Van der Waals contacts between Val21 side chain and the nicotinamide ring,
nicotinamine ribose, and phosphate oxygen atoms are missing in the INH-resistant I21V
InhA structure as compared to WT-sensitive InhA. These interactions do not occur due
to the absence of the CD1 atom in the mutated residue Val21 (Oliveira et al., 2006),
which appear to play an important role in stabilizing bound NADH or INH-NAD in the
InhA active site (Oliveira et al., 2006; Basso et al., 1998).
The crystal structure of S94A InhA showed that a conserved hydrogen bond to a
water molecule is lost owing to mutation to Ala94 (Oliveira et al., 2006), which has
been proposed to account for the reduction in affinity for NADH observed for the INH-
resistant S94A InhA enzyme (Basso et al., 1998). Vilchèze et al. (2006) have solved the
structure of INH-resistant S94A InhA enzyme (PDB code 2NV6) and INH sensitive
InhA in complex with INH-NAD (data do not deposited in the PDB). These authors
propose the loss of the serine causes a shift in position of water molecule that promotes
a disruption in hydrogen binding network. Analysis of the structure of S94A InhA
enzyme present here does not present the disruption between this water molecule and
the atom O9 of the molecule of INH-NAD. Analysis of other structures of S94A InhA,
such a ones solved by Oliveira et al. (2006), shows a binding between this conserved
84
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
12
water molecule and the O9 of the molecule of INH-NAD. Thus, we believed that the
mutation S94A can cause increase flexibility or decrease affinity of this molecule due
the loss binding between this water molecule and the OG atom of the Ser94, which is
observed in the structure wild type. The influence of the mutation on this water
molecule could cause a reduction in the affinity of INH-NAD or NADH by InhA S94A.
Thus, the movement observed in the water molecule by Vilchèze et al. (2006) can be
due an increase of flexibility of this molecule carrying a false impression of an ordered
movement.
The effect of the mutations I21V and S94A on binding process of the NADH on
protein is summarized in the table 2. Analysis of contact area between the InhAs and
INH-NAD reveals that the INH-sensitive WT enzyme presents larger contact area than
the INH-resistant mutants. The value for WT enzyme is 467.4Å
2
while for S94A and
I21V mutants the values are 464.2 Å
2
and 461.9 Å
2
, respectively. The smaller value
observed for I21V mutant can be due the absence of the contact between the CD1 of
Ile21 that is missing in the I21V mutant. The I21V and S94A mutations do not seem to
alter the position of the isonicotinic-acyl group because there are no significant changes
when the mutan t structures are compared with WT type structure (figure 3A-C). These
results are in agreement with the proposal that INH resistance is due to reduction in
NADH affinity for mutant InhAs (S94A and I21V) thereby hampering binding of INH-
NAD adduct.
Effect of INH-NAD on the binding of substrate in the InhA
The isonicotinic-acyl group of INH-NAD adduct replaces the 4S hydrogen of C4
of NADH, which is the same position involved in the hydride transfer catalyzed by
InhA. As a member of the family of SDR enzymes, the catalytic triad formed by
85
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
13
Phe149, Tyr158 and Lys165 (Rozwarski et al., 1999) is present in InhA (Figure 4). The
Phe149 may play a role in anchoring the nicotinamide moiety of NADH for hydride
transfer to the fatty acyl substrate. Lys165 interacts with the 3’-hydroxyl oxygen of the
nicotinamide ribose of NADH, in agreement with a previous proposal that Lys165 plays
a role in cofactor binding (Parikh et al., 1999).
The proposed mechanism of how InhA catalyzes the reduction of the 2-trans
double bond of the fatty acyl substrate consists of the formation of an enolate
intermediate through the direct transfer of a hydride ion from NADH to position C3 of
the substrate, followed by protonation of position C2 (Rozwarski et al., 1999). It is
noteworthy that the Tyr158 occupies approximately the same position for S94A:INH-
NAD and apo S94A, whereas there is a conformational change upon binding of NADH
to S94A, and NAD
+
and a C16 enoyl substrate to WT InhA resulting in position of
Tyr158 to similar positions in these latter structures (figure 4). A comparison between
the WT InhA-NADH structure and an inactive ternary complex formed by WT InhA,
NAD
+
, and a C16 enoyl substrate revealed that upon binding the enoyl substrate there is
a 60° rotation about the Tyr158 Cα-Cβ bond that enables the Tyr158 to hydrogen bond
to the substrate carbonyl group (Parikh et al., 1999). It has thus been suggested that
Tyr158 provides electrophilic stabilization of the transition state(s) for the reaction by
hydrogen bonding to the carbonyl of the substrate. However, the InhA:NAD:C16 enoyl
substrate is a non-productive ternary complex and is likely not part of the reaction
course, which is in agreement with the structural results presented here for apo S94A
InhA. The isonicotyl-acyl group causes a large change in the position of the Phe149 side
chain (figure 4), and appears to be involved in stacking interactions with the INH-NAD
adduct.
86
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
14
INH-resistant S94A mutant in an uncomplexed form
The structure of unliganded S94A InhA reveals that the protein undergoes three
noticeable conformational changes in its main chain upon NADH binding. The main
alterations occur between the residues 99 and 112, between the residues 197 and 213
(figure 5A), and in the Phe41 (figure 5B). It appears that NADH binding is sufficient to
cause a closure of the active site of InhA. Analysis of B-factor for three different
structures of InhA (figure 6) shows that in the absence of both substrate and cofactor or
in the presence of ones, the protein presents the substrate binding loop (residues 196-
219) more disordered than the stru cture in the presence of cofactor. By analysis of these
three structure, we can observe that in the uncomplexed structure, part of the substrate
binding loop is closed (197-203) and part is opened (205-214) for InhA:NAD
+
:substrate
structure (figure5A). In the S94A:NADH structure, part of the substrate binding loop is
more open (197-203) whereas part is more closed (205-214). The conformational
change of InhA:NAD
+
:substrate ternary complex is due to the geometry of the α-helix
of the longer substrate binding loops that could sterically hinder the loop from folding
downward into the substrate binding cavity, leaving a larger opening for the fatty acyl
substrate between residues 205-214 (Rozwarski et al., 1999). The residues 205-214 of
the substrate binding loop moves approximately 6 Å away from the active site in the
unliganded InhA, while the substrate of InhA:NAD
+
:substrate presents the same loops
away approximately . The opening of the substrate binding loop for these structures
are consistent with apoenzyme providing access to substrates to enter the enzyme active
site and non-productive ternary complex providing access to solvent for product release.
Analysis of molecular surface shows the effect of the substrate binding loop on the
active site of InhA (figure 7).
87
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
15
In this manner, the binding of NADH in the active site of the protein can cause
the partial closure of substrate binding loop however arranging a conformation for the
binding of the substrate. In the presence of the co-factor and substrate the protein
presents the active site newly opened for liberation of the product of the reaction. Thus,
in accord ing with th is hypothesis the molecule of NADH is the main responsible by
closure of the active site of InhA.
The van der Waals interactions between the Phe41 and NADH molecule may
play in NADH binding to the enzyme active site, since Phe41 moves closer to the
adenine moiety upon cofactor binding (figure 5B). Asp42 and Arg43 side chains also
undergo conformational changes upon binding of NADH or both NADH and substrate
(figure 5B). However, the absence of substrate does not seem to alter the oligomeric
state of InhA because the tetrameric form in the asymmetric unit and the contact area
among the monomers in the unliganded structure are similar to the other structures
(InhA:NAD
+
:substrate ternary complex and InhA:NADH binary complex) (data not
show). Thus the movement of substrate binding loop appears not to interfere with the
protein oligomeric state.
Conclusion
Here we present four structures of InhA from M. tuberculosis: three structures
for complexes of InhA:INH-NAD and one in the unliganded form. The structures in
complex with INH-NAD are: INH-sensitive WT and two INH-resistant mutants (I21V
and S94A). Comparison between our WT InhA:INH-NAD structure and the structure
previously determined by Roswarski et al., (1998)
reveals that there are changes in
important residues in the active site that were not previously observed. Moreover, the
comparison of our structures in complex with INH-NAD shows that there is no large
88
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
16
influence of mutation in the binding of INH-NAD. The INH-NAD adduct has been
shown to be a slow, tight-binding competitive inhibitor of WT and INH-resistant InhA
mutant enzymes (Rawat et al., 2003). Interestingly, the kinetic and thermodynamic
parameters for the interaction of isonicotinyl-NAD
+
adduct with INH-resistant I21V,
I47T, and S94A InhA mutant enzymes were found to be similar to those of the WT
enzyme (Rawat et al., 2003). These results prompted the authors to suggest an
alternative hypothesis to explain for INH resistance mechanism in strains harboring
inhA-structural gene mutations, in which Inh A may interact directly with o ther
components of the FAS-II system. Accordingly, the resistance-associated mutations in
the inhA-structural gene would affect the susceptibility of InhA to INH inhibition only
in the context of the multienzyme complex, and not when InhA is tested in isolation as
in in vitro assays. Several protein-protein interactions between FAS-II enzymes have
been detected by yeast two-hybrid and co-immunoprecipitation studies and proposed
that either these complexes might coexist or the quaternary structure of aunique” FAS-
II might change from one composition to another during the time and according to the
degree of elongation of the substrate (Veyron-Churlet et al., 2004). In particular, M.
tuberculosis InhA was shown to interact with KasA (β-ketoacyl synthase A) and this
protein-protein interaction has been suggested as a probable explanation to occurrence
of INH-resistant mutant in KasA, even if InhA is indeed the only primary target of INH.
However, it remains to be shown whether inhA structural gene mutations identified in
INH-clinical isolates of M. tuberculosis will affect the inhibition of InhA by INH in the
context of, for instance, InhA-KasA multienzyme complex. In agreement with InhA as
the primary target for INH mode of action, recessive mutations in M. smegmatis and M.
bovis BCG ndh gene, which codes for a type II NADH dehydrogenase (NdhII), have
been found to increase intracellular NADH/NAD ratios (Vilchèze et al., 2005).
89
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
17
Increasing NADH levels protected InhA against inhibition by the INH-NAD adduct
formed upon KatG activation of INH. Hence, mutations in mycobacterial ndh gene
resulted in increased intracellular NADH concentrations, which competitively inhibits
the binding of INH-NAD adduct to InhA, in agreement with the higher dissociation
constant values for NADH found for INH-resistant clinical isolates harbouring inh A-
structural gene mutations as compared to WT InhA (Basso et al., 1998). Moreover, it
has been shown that subtle structural changes result in increased values for the limiting
dissociation rate constant for NADH from INH-resistant mutants (Oliveira et al., 2006).
These observations are in agreement with the results described here.
The unliganded structure presented here shows conformational changes that
occur upon binding of substrates (NADH and/or fatty acyl substrate). The structure
presents movements in the substrate binding loop and a striking change in the position
of Phe41, which may play an important role in NADH stabilization in the enzyme active
site. NADH or substrate binding causes no apparent change in the oligomeric state of
InhA.
The results presented here provide structural information that should better our
understanding of the mechanism of action of, and resistance to, isoniazid in M.
tuberculosis. The unliganded structure of InhA allows identification of conformational
changes upon ligand binding that are important for substrate anchoring. We hope that
these results will help structure-based drug design of more potent antimycobacterial
agents.
PDB Accession codes:
Protein Data Bank: Atomic coordinates and structure factors have been deposited with
accession codes: 2idz, 2ie0, 2ieb, 2ied.
90
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
18
Acknowledgments: Financial support for this work was provided by Millennium
Initiative Program MCT-CNPq, Ministry of Health-Department of Science and
Technology (DECIT)-UNESCO (Brazil), and PRONEX/CNPq/FAPERGS (Brazil) to
D.S.S. and L.A.B. This work was also supported by grants from FAPESP (SMOLBNet,
proc. 01/07532-0, 03/12472-2, 04/00217-0) to WFA. D.S.S. (304051/1975-06), L.A.B.
(520182/99-5), and WFA (CNPq, 300851/98-7) are research career awardees from
National Research Council of Brazil (CNPq).
References
Banerjee, A., Dubnau, E., Quémard, A., Balasubramanian, V., Um, K.S., Wilson, T.,
Collins, D., de Lisle, G., Jacobs, W.R. Jr., 1994. inhA, a gene encoding a target for
isoniazid and ethionamide in Mycoba cterium tuberculosis. Science 263, 227-230.
Basso, L.A., Blanchard, J.S., 1998. Resistance to antitubercular drugs. Adv. Exp. Med.
Biol. 456, 115-144.
Basso, L.A., Santos, D.S., 2005. Drugs that inhibit mycolic acid biosynthesis in
Mycobacterium tuberculosis an update. Med. Chem. Ver. – online 2, 393-413.
Basso, L.A., Zheng, R., Blanchard, J.S., 1996. Kinetics of inactivation of WT and
C243S mutant of Mycobacterium tuberculosis enoyl reductase by activated
isoniazid. J. Am. Chem. Soc. 118, 11301-11302.
91
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
19
Basso, L.A., Zheng, R., Musser, J.M., Jacobs, W.R. Jr., Blanchard. J.S., 1998.
Mechanisms of isoniazid resistan ce in Mycobacterium tuberculosis: enzymatic
characterization of enoyl reductase mutants identified in isoniazid-resistant clinical
isolates. J. Infect. Dis. 178, 769–775.
Bernstein, J.W., Lott, A., Steinberg, B.A., Yale, H.L., 1952 Chemotherapy of
experimental tuberculosis. V. Isonicotinic acid hydrazide (nydrazid) and related
compounds. Am. Rev. Tuberc. 65, 357-374.
Blanchard, J.S., 1996. Mo lecular mechanisms of dru g resistan ce in Mycobacterium
tuberculosis. Ann. Rev. Biochem. 65, 215-239.
CDC (Centers for Disease Control and Prevention), 2006. Emergence of
Mycobacterium tuberculosis with extensive resistance to second-line drugs –
worldwide, 2000-2004. Morb. Mortal Wkly. Rep. 55, 301-305.
Collaborative Computational Project, Number 4., 1994. The CCP4 suite: programs for
protein. Acta Crystallog. Sect. D 50, 760763.
Corbett, E.L., Watt, C.J., Walker, N., Maher, D., Williams, B.G., Raviglione, M.C.,
Dye, C., 2003. The growing burden of tuberculosis: global trends and interactions
with the HIV epidemic. Arch. Intern. Med. 163, 1009-1021.
DeLano, W.L., 2004. The PyMOL Molecular Graphics System. CA DeLano Scientific.
92
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
20
Dessen, A., Quémard, A., Blanchard, J.S., Jacobs, W.R. Jr., Sacchettini, J.C., 1995
Crystal structure and function of the isoniazid target of Mycobacterium tuberculosis.
Science 267, 16381641.
Espinal, M.A., 2003. The global situation of MDR-TB. Tuberculosis: 83, 44-51.
Glickman, M.S., Jacobs, W. R. Jr., 2001 Microbial pathogenesis of Mycobacterium
tuberculosis: dawn of a discipline. Cell, 104, 477-485.
Guex, N., Peitsch, M.C., 1997. SWISS-MODEL and the Swiss-PdbViewer: An
environment for comparative protein modeling. Electrophoresis 18, 2714-2723.
Johnsson, K., King, D.S., Schultz, P.G., 1995. Studies on the mechanism of action of
isoniazid and ethionamide in the chemotherapy of tuberculosis. J. Am. Chem. Soc.
117, 5009-5010.
Johnsson, K., Schultz, P.G., 1994. Mechanistic studies of the oxidation of isoniazid by
the catalase peroxidase from Mycobacterium tuberculosis. J. Am. Chem. Soc. 116,
7425-7426.
Koradi, R., Billeter, M., Wüthrich, K., 1996. MOLMOL: a program for display and
analysis of macromolecular structures. J. Mol. Graph. 14 (1), 51-55.
Kremer, L., Dover, L.G., Morbidoni, H.R., Vilchèze, C., Maughan, W.N., Baulard, A.,
Tu, S.C., Honore, N., Deretic, V., Sacchettini, J.C., Locht, C., Jacobs, W.R. Jr.,
Besra, G.S., 2003. Inhibition of InhA activity, but not KasA activity, induces
93
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
21
formation of a KasA-containing complex in mycobacteria. J. Biol. Chem. 278,
20547-20554.
Kuo, M.R., Morbidoni, H.R., Alland, D., Sneddon, S,F., Gourlie, B.B., Staveski, M.M.,
Leonard, M., Gregory, J.S., Janjigian, A.D., Yee, C., Musser, J.M., Kreiswirth, B.,
Iwamoto, H., Perozzo, R., Jacobs. W.R. Jr., Sacchettini. J.C., Fidock, D.A., 2003.
Targeting tuberculosis and malaria through inhibition of enoyl reductase:
Compound activity and structural data. J. Biol. Chem. 278(23), 20851-20859.
Larsen, M.H., Vilchèze, C., Kremer, L., Besra, G.S., Parsons, L., Salfinger, M., Heifets,
L., Hazbon, M.H., Alland, D., Sacchettini, J.C., Jacobs, W.R. Jr., 2002.
Overexpression of inhA, but not kasA, confers resistance to isoniazid and
ethionamide in Mycobacterium smegmatis, M. bovis BCG and M. tuberculosis. Mol.
Microbiol. 46, 453-466.
Laskowski, R.A., MacArthur, M.W., Moss, D.S., Thornton, J.M., 1993. PROCHECK:
a program to check the stereochemical quality of protein structures. J. Appl.
Crystallog. 26, 283291.
Lei, B., Wei, C.-J., Tu, S.-C., 2000. Action mechanism of antitubercular isoniazid. J.
Biol. Chem. 275, 2520-2526.
McRee, D.E., 1999. XtalView/Xfit—a versatile program for manipulating atomic
coordinates and electron density. J. Struct. Biol. 125, 156–165.
94
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
22
Middlebrook, G., Cohn, M.L., 1953. Some observations on the pathogenicity of
isoniazid-resistant variants of tubercle bacilli. Science 118, 297-299.
Mitchison, D.A., 1985. The action of antituberculosis drugs in short-course
chemotherapy. Tubercle 66, 219-225.
Morlock, G.P., Metchock, B., Sikes, D., Crawford, J.T., Cooksey, R.C., 2003. ethA,
inhA, and katG loci of ethionamide-resistant clinical Mycobacterium tuberculosis
isolates. Antimicrob. Agents Chemother. 47, 37993805.
Murshudov, G.N., Vagin. A.A., Dodson. E.J., 1997. Refinement of macromolecular
structures by the maximum-likelihood method. Acta Crystallog. sect. D 53, 240–
255.
Navaza, J., 1994. AMoRe: an automated package for molecular replacement. Acta
Crystallog. Sect. A 50. 157–163.
No authors listed, 2006 Addressing the threat of tuberculosis caused by extensively
drug-resistant Mycobacterium tuberculosis. Wkly Epidemiol. Rec. 21, 386-390.
Oliveira, J.S., Pereira, J.H., Canduri, F., Rodrigues, N.C., de Souza, O.N., de Azevedo,
W.F. Jr., Basso, L.A., Santos, D.S., 2006. Crystallographic and pre-steady-state
kinetics studies on binding of NADH to wild-type and isoniazid-resistant enoyl-
ACP(CoA) reductase enzymes from Mycobacterium tuberculosis. J. Mol. Biol.
359(3), 646-66.
95
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
23
Oliveira, J. S. Vasconcelos, I. B., Moreira, I. S., Santos, D. S., Basso, L. A., 2007.
Enoyl reductases as targets for the development of anti-tubercular and antimalarial
agents. Curr. Drug Targets 8(3), 399-411.
Ormerod, L.P., 2005. Multidrug-resistant tuberculosis (MDR-TB): epidemiology,
prevention and treatment. British Medical Bulletin 73-74(1), 17-24.
Pablos-Mendez, A., Gowda, D.K., Frieden, T.R., 2002. Controlling multidrug-resistant
tuberculosis and access to expensive drugs: a rational framework. Bull World Health
Organ 80, 489-495.
Parikh, S., Moynihan, D.P., Xiao, G., Tonge, P.J., 1999. Roles of tyrosine 158 and
lysine 165 in th e catalytic mechanism of InhA, the enoyl-ACP reductase from
Mycobacterium tuberculosis. Biochemistry 38, 13623-13634.
Polikarpov, I., Perles, L.A., de Oliveira, R.T., Oliva, G., Castellano, E.E., Garratt, R.C.,
Craievich, A., 1998. Set-up and experimental parameters of the protein
crystallography beam line at the Brazilian National Synchrotron Laboratory. J.
Synchrotron Radiat. 5, 72–76.
Quémard, A., Sacchettini, J.C., Dessen, A., Vilcheze, C., Bittman, R., Jacobs, W.R. Jr.,
Blanchard, J.S., 1995. Enzymatic characterization of the target for isoniazid in
Mycobacterium tuberculosis. Biochemistry 34, 8235-8241.
96
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
24
Rawat, R., Whitty, A., Tonge, P., 2003. The isoniazidNAD adduct is a slow, tight-
binding inhibitor of InhA, the Mycobacterium tuberculosis enoyl reductase: adduct
affinity and drug resistance. Proc. Natl. Acad. Sci. USA 100, 1388113886.
Rozwarski, D.A., Grant, G.A., Barton, D.H.R., Jacobs, W.R., Sacchettini, J.C., 1998.
Modification of the NADH of the isoniazid target (InhA) from Mycobacterium
tuberculosis. Science 279, 98-102.
Rozwarski, D.A., Vilchèze, C., Sugantino, M., Bittman, R., Sacchettini, J.C., 1999
Crystal structure of the Mycobacterium tuberculosis enoyl-ACP reductase, InhA, in
complex with NAD
+
and a C16 fatty acyl substrate. The Journal of Biological
Chemistry 274(22), 15582-15589.
Schroeder, E.K., de Souza, O.N., Santos, D.S., Blanchard, J.S., Basso, L.A., 2002 Drugs
that inhibit mycolic acid biosynthesis in Mycobacterium tuberculosis. Curr. Pharm.
Biotechnol. 3, 197–225.
Uchoa, H.B.; Jorge, G.E.; Freitas Da Silveira, N.J.; Camera, J.C.,Jr.; Canduri, F.; De
Azevedo, W.F.,Jr. Biochem. Biophys. Res. Commun., 2004, 325, 1481-1486.
Veyron-Churlet, R., Guerrini, O., Mourey, L., Daffe, M., Zerbib, D., 2004. Protein-
protein interaction s within the fatty acid synthase-II system of Mycobacterium
tuberculosis are essential for mycobacterial viability. Mol. Microbiol. 54, 1161-
1172.
97
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
25
Vilchèze, C., Morbidoni, H.R., Weisbrod, T.R., Iwamoto, H., Kuo, M., Sacchettini,
J.C., Jacobs, W.R. Jr., 2000. Inactivation of the inhA-encoded fatty acid synthase II
(FASII) enoyl-acyl carrier protein reductase induces accumulation of the FASI end
products and cell lysis of Mycobacterium smegmatis. J. Bacteriol. 182, 4059-4067.
Vilchèze, C., Wang, F., Arai, M., Hazbon, M.H., Colangeli, R., Kremer, L., Weisbrod,
T.R., Alland, D., Sacchettini, J.C., Jacobs, W.R. Jr., 2006. Transfer of a point
mutation in Mycobacterium tuberculosis inhA resolves the target of isoniazid. Nat
Med: 12, 1027-1029.
Vilchèze, C., Weisbrod, T.R., Chen, B., Kremer, L., Hazbon, M.H., Wang, F., Alland,
D., Sacchettini, J.C., Jacobs, W.R. Jr., 2005. Altered NADH/NAD
+
ratio mediates
coresistance to isoniazid and ethionamide in mycobacteria. Antimicrob. Agents
Chemother. 49, 708-720.
Zabinski, R.F., Blanchard, J.S., 1997. The requirement for manganese and oxygen in the
isoniazid-dependent inactivation of Mycobacterium tuberculosis enoyl reductase. J.
Am. Soc. Chem. 119, 2331-2332.
98
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
26
Figures
Figure 1. Ribbon diagram of InhA. A) Monomeric structure of the wild type InhA:INH-
NAD binary complex, and B) Tetrameric structure of the unliganded InhA. The figures
were drawn using program MolMol (Koradi et al., 1996).
Figure 2. Superposition of active site residues of InhA. A) It is shown the superposition
of the active site structure of wild type InhA: INH-NAD solved by Roswarski et al.,
(1998) at 2.7Å resolution (in green) and wild type InhA: INH-NAD at 2.0 Å resolution
presented here (in grey). B) It is shown the superposition of the structure of wild type
InhA:NADH complex binary solved by Oliveira et al., (2006) (in blue) and the
structure of wild type InhA: INH-NAD presented here (in grey). The figures were
generated with MolMol (Koradi et al., 1996).
Figure 3. Ribbon diagrams for InhA structures, showing the electronic density for INH-
NAD and mutated residue. A) Wild type InhA, B) I21V mutant and C) S94A mutant.
The figures were generated using program SpdbView (Guex and Peitsch, 1997).
99
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
27
Figure 4. Catalytic triad for InhA. In green is shown the residues of S94A InhA in
complex with INH-NAD, in grey is shown the residues of S94A InhA unliganded, in
orange is shown the residues of S94A in complex with NADH, and in blue is shown the
residues of WT InhA in complex with C16 enoyl substrate and NAD
+
(Rozwarski
et al.,
1999. The Figure shows conformational changes in these residues due to different
ligands. The figure was generated using program Molmol (Koradi et al., 1996).
Figure 5. Conformational changes observed in the structure of S94A InhA upon
binding of NADH or both NADH and substrate. A) Cα traces of S94A InhA unliganded
(purple), S94AInhA in complex with NADH (orange), and WT InhA in complex with
both NAD
+
and substrate (green). B) Conformational changes of residues 41-43
(coloring was the same as in A). The figures were generated using program MolMol
(Koradi et al., 1996).
Figure 6. Residue-averaged B factor for S94A resistant mutant apoenzyme (red line),
WT InhA in complex with NAD
+
and C16 substrate (black line), and S94A resistant
mutant in complex with NADH (purple line).
Figure 7. Molecular surface for A) Unliganded S94A InhA, B) S94A InhA in complex
with NADH, and C) WT InhA in complex with both NADH and substrate. The figures
were generated using program Pymol (DeLano, 2004).
100
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Table 1. Crystallographic data, refinement statistics and analysis of the quality of InhA structures
Crystallographic data
WtInhA:NADH:INH I21V InhA:NADH-INH S94A InhA:NADH-INH S94A InhA
a (Å) 97.09 96.59 96.83 54.63
b (Å) 97.09 96.59 96.83 63.52
c (Å) 136.94 136.35 136.42 65.18
α
90 90 90 97.21
β
90 90 90 85.81
γ
120 120 120 102.87
Space group P6222 P6222 P6222 P1
Number of image 80 60 80 160
Number of measurements with I>2 σ (I)
245823 134890 185407 77775
Number of independent reflections 26409 19563 19782 44054
Completeness (%) (Outermost shell) 99.9 (100) 98.7 (91.6) 99.8 (99.5) 98.6 (85.2)
R
*
sym
(%)
a
(Outermost shell) 9.8 (83.0) 7.2 (59.1) 9.5 (77.4) 5.8 (29)
Redundancy 9.3 6.9 9.4 1.6
Refinement statistics
Resolution range (Å) 40.19-2.0 39.41-2.2 45.64-2.2 42.26-2.14
Reflections used for refinement 25045 18527 18193 41818
Number of atoms 2283 2261 2257 8682
Number of residues 267 267 267 1068
Number of water molecules 239 216 212 701
Final R-factor (¨%)
b
18.7 18.4 17.9 16.2
Final R-factor free (%)
c
22.6 23.6 23.4 25.5
Correlation Coefficient (%) 96.0 95.8 95.8 95.8
B values (A
2
)
Main chain 30.56 36.69 34.08 27.52
Side chain 32.19 37.70 35.33 28.59
NADH-INH 32.19 31.93 29.85
Waters 40.93 43.39 42.13 33.23
Table1_rev
101
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Ramachandran plot
Favorable 86.2 87.1 88.9 88
Additional allowed 9.8 9.8 9.3 10.3
Generously allowed 0.0 0.0 0.0 0.7
Disallowed 4.0 3.1 1.8 1.0
a
R
sy m
= 100 P I (h) _ Æ I (h/P I (h) with I (h), observed intensity and ÆI (h)æ, mean intensity of reflection h overall measurement of I (h).
b
R-factor = 100 × ( F
obs
F
calc
)/ (F
obs
), the sums being taken over all reflections with F/σ(F) > 2σ(F).
c
R -free = R-factor for 10% of the data that were not included during crystallographic refinement.
102
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Table 2. Effect of the mutations I21V and S94A on the binding process of the NADH on
the InhA.
mutation Mutation effect consequence
I21V
Lose of Van der Waals
interaction between NADH
and the CD1 atom present in
the valine residue
Decrease of the stability of
the binding of the NADH in
the active site of the protein
S94A
Alteration in the binding
network involving a
conserved water molecule
and O9 atom of molecule of
NADH
Increase of the flexibility of
the conserved water
molecule and decrease of
the affinity of NADH by
protein
Table2_rev
103
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure1a_rev
Click here to download high resolution image
104
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure1b_rev
Click here to download high resolution image
105
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure2a_rev
Click here to download high resolution image
106
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure2b_rev
Click here to download high resolution image
107
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure3a_rev
Click here to download high resolution image
108
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure3c_rev
Click here to download high resolution image
109
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure3b_rev
Click here to download high resolution image
110
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure4_rev
Click here to download high resolution image
111
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure5a_rev
Click here to download high resolution image
112
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure5b_rev
Click here to download high resolution image
113
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure6_rev
Click here to download high resolution image
114
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure7a_rev
Click here to download high resolution image
115
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure7b_rev
Click here to download high resolution image
116
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Figure7c_rev
Click here to download high resolution image
117
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
4. CONCLUSÃO
Este trabalhou teve como principal objetivo análise estrutural de enzimas que são
importantes para a viabilidade de M. tuberculosis, mas que, entretanto tais enzimas
estivessem ausentes em humanos. Assim essas enzimas são consideradas como alvos para
o desenvolvimento de novas drogas contra a tuberculose e podem apresentar baixa
toxidade ou ausência de efeitos colaterais. Após os estudos realizados com cada uma das
enzimas chegamos as seguintes conclusões para cada uma delas separadamente:
- Foi determinada a primeira condição de cristalização para a proteína corismato
sintase de M. tuberculosis, na qual foram obtidos cristais hexagonais, grupo
espacial P6
4
22. Estes cristais difrataram em torno de 2,65Å de resolução e os dados
do processamento de difração de raios X foram utilizados para a resolução da
estrutura cristalográfica desta proteína;
- Foi determinada a estrutura cristalográfica da corismato sintase de Mycobacterium
tuberculosis em seu estado nativo a 2,65Å de resolução. A estrutura apresenta um
folding bastante similar às outras estruturas de corismato sintase de outros
microorganismos. Os resíduos do sítio ativo tamm são bastante conversados,
entretanto a posição de alguns e resíduos apresenta diferente conformações quando
comparado com outras estruturas. Essas diferenças podem ser devido à ausência de
ligantes. A determinação de estruturas da corismato sintase de M. tuberculosis em
complexo com seus substratos podem esclarecer essas evidências;
- Foram construídos seis modelos estruturais para a proteína triptofano sintase de M.
tuberculosis com diferentes inibidores análogos de substratos através da técnica de
modelagem molecular comparativa. Os modelos obtidos apresentam qualidade
estereoquímica satisfatória e mostram as interações entre estes ligantes e a enzima.
118
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Pela análise da ligação podemos concluir que os inibidores estudados apresentam
uma interação satisfatória com a enzima. Testes in vitro ou in vivo poderão
confirmar essa hipótese. Entretanto os modelos obtidos podem ser utilizados na
construção de outros inibidores mais potentes para esta enzima.
- Foram apresentadas também duas estruturas da proteína chiquimato quinase de
Mycobacterium tuberculosis (uma em complexo com ADP e chiquimato e outra em
complexo com MgADP). A estrutura da chiquimato quinase em complexo com
ADP e chiquimato mostra que a ausência do íon magnésio pode influenciar a
posição dos grupos hidroxilas da molécula de chiquimato e também em alguns
importantes resíduos do sítio ativo da enzima, como o Asp34. A estrutura da
chiquimato quinase em complexo com MgADP foi obtida em uma nova condição
de cristalização e apresenta-se como um tetrâmero na unidade assimétrica, com
simetria 222. Nesta estrutura pôde-se observar o possível efeito do chiquimato
sobre o empacotamento cristalino da enzima. Além disso, devido à disposição do
monômero na unidade assimétrica foi possível observar que a molécula ADP tem
uma importante participação na estabilidade da estrutura quaternária da proteína,
pelo menos na rede cristalina. Um dos monômeros do tetrâmero observado para a
chiquimato quinase não apresentava o íon cloreto e assim é possível que ele tenha
uma importante participação da orientação da molécula de ADP no sítio ativo da
enzima, no fechamento do LID domain e também na formação de interações com
moléculas de águas que podem levar o fechamento da enzima.. Desta forma, este
íon pode ter importante influência na estabilidade e no mecanismo de catálise da
enzima. Assim, as informações obtidas neste trabalho podem ser úteis para o
entendimento do mecanismo catalítico desta enzima e também para o desenho de
novos ligantes que tenham atividade inibitória.
119
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
- Foram apresentadas as estruturas cristalográficas da proteína Inha selvagem e de
dois mutantes isolados clínicos resistentes a isoniazida (I21V e S94A) complexada
com o aducto NAD-INH. A partir destas estruturas é possível determinar as
alterações moleculares que a molécula de NAD-INH realiza no sítio ativo da
enzima e também como ocorre o efeito inibitório desta droga. Foi observada nas
estruturas dos mutantes em complexo com NAD-INH que, as mutações I21V e
S94A parecem não influenciar na interação entre a proteína e o aducto. E assim,
estas informações corroboram com os dados existentes na literatura de que a
resistência pode estar relacionada com uma menor afinidade pela molécula de
NADH. Entretanto a resistência a isoniazida poderia estar associada também a
outros fatores nas quais estas mutações poderiam de forma indireta contribuir. É
apresentado também neste trabalho a estrutura do mutante de InhA S94A na sua
forma nativa, em um novo grupo espacial ainda não descrito na literatura para esta
enzima (P1). Nesta estrutura puderam-se concluir quais são os possíveis
movimentos que podem ocorrer durante o processo de ligação do NADH e do
substrato acil C16. Além disso, foi possível evidenciar importantes resíduos que
podem contribuir para a estabilidade da ligação entre o substrato e a enzima e
importantes movimentos que ocorrem em alguns resíduos do sítio ativo durante o
processo de catálise. Esperamos que estas informações contribuam para o
entendimento do mecanismo de inibição e de resistência à isoniazida e que possam
ser úteis na construção de novas drogas contra tuberculose.
120
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
5. BIBLIOGRAFIA
AHN, H. J. et al. Crystal structure of chorismate synthase: a novel FMN-binding protein
fold and functional insights. J. Mol. Biol. v. 336(4), p 903-915, 2004.
BAPTISTA, I. M. et al. Drug resistance and genotypes of strains of Mycobacterium
tuberculosis isolated from human immunodeficiency virus infected and non-infected
tuberculosis patients in Bauru, São Paulo, Brazil. Mem. Inst. Oswaldo Cruz, v. 97(8), p.
1147-1152, 2002.
BASSO, L. A. et al. Mechanisms of isoniazid resistance in Mycobacterium tuberculosis:
enzymatic characterization of enoyl reductase mutants identified in isoniazid-resistant
clinical isolates. J. Infect. Dis., v. 178, p. 769–775, 1998.
BAUBY, H.; GIRONS, I. S.; PICARDEAU, M. Construction and complementation of the
first auxotrophic mutant in the spirochaete Leptospira meyeri. Microbiology, v. 149, p.
689–693, 2003.
BANNER, D. W. et al. Structure of chicken muscle triose phosphate isomerase determined
crystallographically at 2.5 Å resolution using amino acid sequence data. Nature, v. 255, p.
609-614, 1975.
BENTLEY, R. The shikimate pathway - metabolic tree with many branches.Crit. Rev.
Biochem. Mol. Biol., v. 25, p. 307-384, 1990.
BRENNAN, P. J.; NIKAIDO, P. J. The envelope of mycobacteria. Annu. Rev. Biochem.,
v. 64, p. 29-63, 1995.
CERASOLI, E. et al. Effects of salts on the function and conformational stability of
shikimate kinase. Biochem. Biophys. Acta, v. 1648(1-2), p. 43-54, 2003.
COLE, S. T. et al. Deciphering the biology of Mycobacterium tuberculosis from the
complete genome sequence. Nature, v. 393, p.537-544, 1998.
COTRAN, R. S.; KUMAR, V.; ROBBINS, S. L. Robbins: Patologia estrutural e
funcional. Tradução de Amaury José da Cruz Júnior et al. 4.ed. Rio de Janeiro, Guanabara
Koogan, 1991.
DESSEN, A. et al. Crystal structure and function of the isoniazid target of Mycobacterium
tuberculosis. Science, v. 267, p. 1638–1641, 1995.
DIAS, M. V. B. et al. Crystal structure of chorismate synthase from Mycobacterium
tuberculosis. Journal of Structural Biology, v. 154, p. 130-143, 2006.
DHALIWAL, B. et al. Crystallographic studies of shikimate binding and induced
conformational changes in Mycobacterium tuberculosis shikimate kinase. FEBS Lett. V.
574(1-3), p. 49-54, 2004.
121
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
DOSSELAERE, F.; VANDERLEYDEN, J.A. A metabolic node in action: chorismate-
utilizing enzymes in microorganisms. Crit. Rev. Microbiol., v.27, p.75-131, 2001.
DOVER, L. G. et al. Comparative cell wall core biosynthesis in the mycolated pathogens,
Mycobacterium tuberculosis and Corynebacterium diphtheriae. FEMS Microbiology
Reviews, v. 28(2), p. 225-250, 2004.
DUNCAN, K. Progress in TB drug development and what is still needed. Tuberculosis, v.
83, p. 201-207, 2003.
DYE, C. et al. Erasing the world's slow stain: strategies to beat multidrug-resistant
tuberculosis Science, v. 295, p. 2042-2046, 2002.
FINN, J. et al. Rational herbicide design by inhibition of tryptophan biosynthesis. Bioorg.
Med. Chem. Letters, v. 9, p. 2297-2302, 1999.
GERSTEIN, M.; SCHULZ, G.; CHOTHIA C. Domain closure in adenylate kinase. Joints
on either side of two helices close like neighboring fingers. J. Mol. Biol. v. 229(2), p. 494-
501, 1993.
GAN, J. et al. Crystal structure of Mycobacterium tuberculosis shikimate kinase in complex
with skikimic acid and an ATP analogue. Biochemistry, v. 45, p. 8539-8545, 2006.
GU, Y. et al. Crystal structure of shikimate kinase from Mycobacterium tuberculosis reveals
the dynamic role of the LID domain in catalysis. J. Mol. Biol. v. 319, p. 779-789, 2002.
HARTMANN, M. D. et al. Mechanism of phosphoryl transfer catalyzed by shikimate
kinase from Mycobacterium tuberculosis. J. Mol. Biol., 2006 (in press).
HAWSER, S.; LOCIURO, S.; ISLAM, K. Dihydrofolate reductase inhibitors as
antibacterial agents. Biochemical Pharmacology, v. 71(7), p. 941-948, 2006.
HERRMANN, K. M. The shikimate pathway: early steps in the biosynthesis of aromatic
compounds. The Pant Cell, v. 7, p. 907-919, 1995.
HERRMANN, K. M.; WEAVER, L. M. The shikimate pathway. Annu Rev Plant
Physiol Plant Mol Biol. v. 50, p. 473-503, 1999.
HYDE, C. C. et al., Three-dimensional structure of the tryptophan synthase α
2
β
2
multienzyme com plex from Salmonella typhimurium. J. Bio. Chem. v. 263(33), p. 17857-
17871, 1988.
KHASNOBIS, S.; ESCUYER, V.; CHATTERJEE, D. Emerging therapeutic targets in
tuberculosis: post-genomic era. Expert. Opin. Ther. Targets, v. 6, p. 21-40, 2002.
KING, D. A. et al. Epidemiology. Infectious diseases: preparing for the future. Science, v.
313 (5792), p. 1392-1393, 2006.
KITZING, K. et al., Mechanism of chorismate synthase. Role of the two invariant histidine
residues in the active site. J. Biol. Chem. v. 279(10), p. 9451-9461, 2004.
122
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
KITZING, K.; MACHEROUX, P.; AMRHEIN, N. Spectroscopic and kinetic
characterization of the bifunctional chorismate synthase from Neurospora crassa: evidence
for a common binding site for 5-enolpyruvylshikimate 3-phosphate and NADPH. J. Biol.
Chem.,v. 276, p. 42658-42666, 2001.
KUO, M. R. et al. Targeting tuberculosis and malaria through inhibition of enoyl
reductase: Compound activity and structural data. The journal of Biological Chemistry,
v. 278(23), p. 20851-20859, 2003.
KRELL, T.; COGGINS, J. R.; LAPTHORN, A. J. The three-dimensional structure of
shikimate kinase. J. Mol. Biol. v. 278(5), p. 983-997, 1998.
KUMAR, V.; ABBAS, A. K.; FASTO, N. Robbins e Cotran: patologia – bases
patológicas das doenças. Tradução de Maria da Conceição Zacharias et al. Rio de Janeiro:
Elsevier, 2005. 1592p.
KURADI, R.; BILLETER, M.; WÜTHRICH, K. MOLMOL: a program for display and
analysis of macromolecular structures. J. Mol. Graphics v. 14, p. 51-55, 1996.
MACLEAN J.; ALI S. The structure of chorismate synthase reveals a novel flavin binding
site fundamental to a unique chemical reaction. Structure v. 11(12), p. 1499-511, 2003.
MACHEROUX, P. et al. A unique reaction in a common pathway: mechanism and
function of chorismate synthase in the shikimate pathway. Planta, v. 207(3), 325-334,
1999.
McARTHUR J. D. et al. An aromatic amino acid auxotrophic mutant of Bordetella
bronchiseptica is attenuated and immunogenic in a mouse model of infection. FEMS
Microbiology Letters, v. 221, p. 7-16, 2003.
McREE, D. E. XtalView/Xfit--A versatile program for manipulating atomic coordinates
and electron density. J. Struct. Biol., v. 125, p. 156-165, 1999.
MILES, E. W.; RHEE, S.; DAVIES, D. V. The molecular basis of substrate channeling.
The journal of biological chemistry, v. 274 (18), p. 12193-12196, 1999.
MILES, E. W. Tryptophan synthase. Structure, function, and protein engineering. Subcell
Biochem. v. 24, p. 207-254, 1995.
MOLLE, V. et al. The condensing activities of the Mycobacterium tuberculosis type II
fatty acid synthase are differentially regulated by phosphorylation. J. Bio. Chem., v.
281(40), p. 30094-30103, 2006.
MORGAN, J.; HARITAKUL, R.; KELLER, P. A. Anilinopyrimidines as novel
antituberculosis agents. Bioorg. Med. Chem. Lett., v. 13, p. 1755-1757, 2003.
MÜLLER, C. W. et al. Adenylate kinase motions during catalysis: an energetic
counterweight balancing substrate binding. Structure v. 4, p. 147-156, 1996.
MURSHUDOV, G. N.; VAGIN A. A.; DODSON, E. J. Refinement of Macromolecular
Structures by the Maximum-Likelihood Method Acta Cryst. D53, p. 240-255, 1997.
123
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
OLIVEIRA, J. S. et al. Crystallographic and pre-steady-state kinetics studies on binding of
NADH to wild-type and isoniazid-resistant enoyl-ACP(CoA) reductase enzymes from
Mycobacterium tuberculosis. J. Mol. Biol., v. 359(3), p. 646-66 2006.
PAN, P.; WOEHL, E.; DUNN, M. F. Protein architecture, dinamycs and allostery in
tryptophan synthase channeling. TIBS, v. 22, p. 22-27, 1997.
PARISH, T.; STOKER, N. G. The common aromatic amino acid biosynthesis pathway is
essential in Mycobacterium tuberculosis. Microbiology, v. 148, p. 3069-3077, 2002.
PEREIRA, J. H. et al. Structure of shikimate kinase from Mycobacterium tuberculosis
reveals the binding of shikimic acid. Acta Crystallogr. D. Biol., v. 60 (pt 12 pt 2), p.
2310-2319, 2004.
PELCZAR, M. J. Jr.; CHAN, E. C. S; KRIEG, N. R. Microbiologia – conceitos e
aplicações. Tradução: Sueli F. Yamada et al. 2.ed. São Paulo: Makron Books, 1996. 2 v.
PITTARD, A. J. Biosynthesis of the aromatic amino acids. In Escherichia coli and
Salmonella typhimurium: cellular and molecular biology. Neidhardt, F. C., Washington,
DC: American Society for microbiology, pp 368-394,1987
QUÉMARD, A. et al. Enzymatic characterization of the target for isoniazid in
Mycobacterium tuberculosis. Biochemistry, v. 34, p. 8235–8241, 1995.
QUEVILLON-CHERUEL S. et al. Crystal structure of the bifunctional chorismate
synthase from Saccharomyces cerevisiae. J. Biol. Chem. v. 279(1), p. 619-625, 2004.
RAVIGLIONE, M. C. et al., Assessment of worldwide tuberculosis control. WHO global
surveillance and monitoring project. The lancet, v. 350, p. 624-629, 1997.
RAWAT, R.; WHITTYty, A.; TONGE, P. The isoniazid–NAD adduct is a slow, tight-
binding inhibitor of InhA, the Mycobacterium tuberculosis enoyl reductase: adduct affinity
and drug resistance. Proc. Natl Acad. Sci. USA, v. 100, p. 13881–13886, 2003.
ROBERTS, C. W. et al. The shikimate pathway and its branches in apicomplexan
parasites. J. Infec. Dis., v. 185, Suppl 1, p. S25-36, 2002.
ROBERTS, F. et al. Evidence for the shikimate pathway in apicomplexan parasites.
Nature, v. 393(6687), p. 801-805, 1998
ROMANOWSKI, M. J.; BURLEY, S. K. Crystal structure of the Escherichia coli
shikimate kinase I (AroK) that confers sensitivity to mecillinam. Proteins, v. 47(4), p. 558-
62, 2002.
ROSWARSKI, D. A. et al.
Crystal structure of the Mycobacterium tuberculosis enoyl-
ACP reductase, InhA, in complex with NAD
+
and a C16 fatty acyl substrate. The Journal
of Biological Chemistry, v. 274(22), p. 15582-15589, 1999.
124
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
ROSWARSKI, D. A. et al. Modification of the NADH of the isoniazid target (InhA) from
Mycobacterium tuberculosis. Science, v. 279, p. 98–102, 1998.
SACHPATZIDIS, A. et al. Crystallographic studies of phosphonate-based α-reaction
transition-state analogues complexed to tryptophan synthase. Biochemistry, v. 38, p.
12665-12674, 1999.
SALI, A.; BLUNDELL, T. L. Comparative protein modelling by satisfaction of spatial
restraints. J. Mol. Biol., v. 234, p. 779–815, 1993.
SCHRUEDER, E. K. et al. Drugs that inhibit mycolic acid biosynthesis in Mycobacterium
tuberculosis. Curr. Pharm. Biotechnol., v. 3, p. 197–225, 2002.
SCHULZ, G. E. Binding of nucleotides by proteins. Curr. Opin. Struct. Biol. v. 2, p. 61-
67, 1992.
SKÖLD, O. Resistance to trimethoprim and sulfonamides. Veterinary Research, v. 32, p.
261-273, 2001.
SMITH, D. A. et al. Characterization of auxotrophic mutants of Mycobacterium
tuberculosis and their potential as vaccine candidates. Infection and Immunity, v. 69 (2),
p. 1142-1150, 2001.
TACKET, C. O.; SZTEIN, M. B.; LOSONSKY, G. A.; WASSERMAN, S. S.; NATARO,
J. P.; EDELMAN, R.; PICKARD, D.; DOUGAN, G.; CHATFIELD, S. N.; LEVINE, M.
M. Safety of live oral Salmonella typhi vaccine strains with deletions in htrA and aroC
aroD and immune response in humans. Infect. Immun., v. 65, p. 452-456, 1997.
TROUILLER, P. et al., Drug development for neglected diseases: a deficient market and a
public-health policy failure. Lancet, v. 359(9324), p. 2188-2194, 2002.
TROUILLER, P. et al., Drugs for neglected diseases: a failure of the market and a public
health failure? Tropical Medicine And International Health, v. 6, p.945-951, 2001.
VIOLA, C. M.; SARIDAKIS, V.; CHRISTENDAT, D. Crystal structure of chorismate
synthase from Aquifex aeolicus reveals a novel beta alpha beta sandwich topology.
Proteins, v. 54(1), p. 166-169, 2004.
YAN, H.; YSAI, M. D. Nucleoside monophosphate kinases: structure, mechanism, and
substrat specificity. Advan. Enzymol. Relat. Areas Mol. Biol. v. 73, p. 103-134, 1999.
WANG, R et al. SCORE: a new empirical method for estimating the binding affinity of a
protein-ligand complex. J. Mol. Model. v. 4, p. 379-394, 1998.
WEYAND M.; SCHLICHTING I. Crystal structure of wild-type tryptophan synthase
complexed with the natural substrate indole-3-glycerol phosphate. Biochemistry, v.
38(50), p.16469-16480, 1999.
125
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
4. Anexo
Artigo de Revisão publicado
126
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Chorismate Synthase: an attractive target for drug development against neglected
disease.
Marcio V. B. Dias; Fernanda Ely; Mário S. Palma; Luiz A. Basso; Diógenes S. Santos;
Walter Filgueira de Azevedo Jr. Current Drugs Targets, v. 8, p. 48-55, 2007.
Neste artigo é apresentada uma revisão a respeito das informações sobre a
corismato sintase publicadas em revistas indexadas nos últimos anos.
Nele é descrito a importância desta enzima e da via onde ela se encontra como
alvo para o desenvolvimento de novas drogas. É comentado desde a descoberta do
produto desta enzima, o corismato em 1962 até as informações publicadas em 2004. É
realizado um estudo das características desta enzima em diferentes organismos que a
possui, como fungos, bactérias, plantas e parasitas do filo apicomplexa. Portanto, é
descrita as peculiaridades das corismato sintases monofuncionais encontradas em
bactérias, plantas e parasitas do filo apicomplexa e das bifuncionais encontradas em
fungos. São discutidos também, os dados existentes sobre o possível processo evolutivo
da existência de enzimas monofuncionais e bifuncionais. São mostradas as
peculiaridades das corismato sintases de organismos eurcariotos (Plantas e parasitas de
filo apicomplexa) e sua localização nestas células. É discutido detalhadamente o
mecanismo catalítico desta enzima, com a importância da molécula de FMN para a
transferência de elétrons para a molécula de EPSP e a importância da molécula de
NADPH para as corismato sintases bifuncionais. É realizada uma descrição da estrutura
desta enzima, os aspectos peculiares do seu sítio ativo e a disposição das moléculas de
FMN e EPSP, e como isso pode auxiliar no processo de catálise desta enzima. E
finalmente, são apresentados os estudos já realizados no desenvolvimento de novos
ligantes e as estratégias utilizadas.
127
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Current Drug Targets, 2007, 8, 000-000 1
1389-4501/07 $50.00+.00 © 2007 Bentham Science Publishers Ltd.
Chorismate Synthase: An Attractive Target For Drug Development Against Orphan
Diseases
Marcio V. B. Dias
a*
, Fernanda Ely
b
, Mário S. Palma
c
, Walter F. de Azevedo Jr.
d
, Luiz A. Basso
b
and
Diógenes S. Santos
b**
a
Programa de Pós-Graduação em Biofísica Molecular – Departamento de Física, UNESP, São José do Rio Preto, SP 15054-000,
Brasil;
b
Centro de Pesquisas em Biologia Molecular e Funcional / PUCRS. Avenida Ipiranga 6681, Tecnopuc, Partenon 90619-900,
Porto Alegre, RS, Brazil;
c
Laboratório de Biologia Estrutural e Zooquímica-CEIS / Departamento de Biologia – Instituto de Biociên-
cias, UNESP, Rio Claro, SP 13506-900, Brasil and
d
Faculdade de Biociências-/ PUCRS. Avenida Ipiranga 6681, Partenon 90619-
900, Porto Alegre, RS, Brazil
Abstract: The increase in incidence of infectious diseases worldwide, particularly in developing countries, is worrying. Each year, 14
million people are killed by infectious diseases, mainly HIV/AIDS, respiratory infections, malaria and tuberculosis. Despite the great
burden in the poor countries, drug discovery to treat tropical diseases has come to a standstill. There is no interest by the pharmaceutical
industry in drug development against the major diseases of the poor countries, since the financial return cannot be guaranteed. This has
created an urgent need for new therapeutics to neglected diseases. A possible approach has been the exploitation of the inhibition of
unique targets, vital to the pathogen such as the shikimate pathway enzymes, which are present in bacteria, fungi and apicomplexan para-
sites but are absent in mammals. The chorismate synthase (CS) catalyses the seventh step in this pathway, the conversion of 5-
enolpyruvylshikimate-3-phosphate to chorismate. The strict requirement for a reduced flavin mononucleotide and the anti 1,4 elimination
are both unusual aspects which make CS reaction unique among flavin-dependent enzymes, representing an important target for the che-
motherapeutic agents development. In this review we present the main biochemical features of CS from bacterial and fungal sources and
their difference from the apicomplexan CS. The CS mechanisms proposed are discussed and compared with structural data. The CS struc-
tures of some organisms are compared and their distinct features analyzed. Some known CS inhibitors are presented and the main charac-
teristics are discussed. The structural and kinetics data reviewed here can be useful for the design of inhibitors.
Key Words: Infectious disease; neglected disease; chorismate synthase; shikimate pathway; flavin-dependent enzymes.
INTRODUCTION
A worrying increase in incidence of infectious diseases was recog-
nized in the late 1980’s, causing the suffering of millions of people,
especially in tropical and subtropical areas as Africa, Asia and
South America – which account for four-fifth of the world’s popu-
lation. Each year, 14 million people are killed by infectious dis-
eases, mainly diseases as HIV/AIDS, respiratory infections, malaria
and tuberculosis [1]. The increase of immunodeficient population,
mainly HIV-positive, the large homeless population and decline in
health care structures are some problems responsible for the in-
creased incidence of many infectious diseases in the world. Among
the neglected infectious diseases are tuberculosis and malaria. The
Mycobacterium tuberculosis, the aetiological agent of tuberculosis
(TB), kills more than 3 million people each year and ninety percent
of TB cases occur in developing countries
[2]. In addition, malaria,
caused by apicomplexan parasites of the genus Plasmodium, infects
millions of people and account for the death of more than 2 million
children annually [3]. In contrast, despite the enormous burden in
the poor countries, drug discovery and development targeted at
tropical diseases are at a standstill [1, 4]. Today, the pharmaceutical
industry is reluctant to invest in drug development to treat the major
diseases of the poor countries, because the financial return cannot
be guaranteed. Thus, an urgent reorientation of priorities in drug
development and health policy is needed, principally by national
and international policies that need to direct the global economy
and address the true health needs of society [1].
With the completion of the genome sequences of several patho-
genic organisms is occurring an enormous impact on our under-
*Address correspondence to this author at the Programa de Pós-Graduação
em Biofísica Molecular – Departamento de Física, UNESP, São José do Rio
Preto, SP 15054-000, Brasil; Tel: ---------------------; Fax: -------------------;
E-mail: diogenes@pucrs.br
standing of the pathogenicity of these organisms. Thus, the infor-
mation obtained by genome projects can be used by national and
international research policy makers and scientific community for
development of drugs against many of these diseases. The genome
allows identification of metabolic pathways present in the patho-
genic microorganisms, which can be target for the development of
new drugs. A possible approach to selective antimicrobial chemo-
therapy has been to exploit the inhibition of unique targets, vital to
the pathogen and absent in mammals [5]. The shikimate pathway is
an attractive example of this kind of targets, since it is present in
bacteria, fungi and apicomplexan parasites but absent from mam-
mals [6]. In this pathway the glycolytic intermediate, phosphoenol
pyruvate, and the pentose phosphate pathway intermediate, D-
erytrose 4-phosphate, are converted to chorismate through seven
metabolic steps [7]. The essentiality of shikimate pathway was ob-
served in some microorganisms such as Plasmodium falciparum
and M. tuberculosis. The disruption of aroK gene, which codes for
the shikimate kinase, showed that this enzyme is essential for M.
tuberculosis viability [8]. In P. falciparum the growth was inhibited
by glyphosate, a well-characterized inhibitor of the shikimate path-
way [9]. These reports provide strong evidence that shikimate
pathway is essential for the survival of these pathogens; therefore,
its enzymes are potential targets for drug development.
The product of shikimate pathway, the chorismate or chorismic
acid, is a dihydroaromatic compound and it was first described by
Frank and Margaret Gibson in 1962 [10]. This compound is the
branch point in the biosynthesis of several important aromatic
molecules. For this reason, it was named chorismate, which means,
in Greek, separation, split, or divorce. The chorismate is the com-
mon precursor for the biosynthesis of a wide range of primary and
secondary metabolites including aromatic amino acids (phenyla-
lanine, tyrosine and tryptophan), folate, naphthoquinones, menaqui-
nones and mycobactins [11]. Chorismate synthase (CS), which is
responsible for the synthesis of chorismate, was first described in
1967 by Morell et al. [12]. This enzyme catalyses the seventh step
128
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
2 Current Drug Targets, 2007, Vol. 8, No. 3 Dias et al.
in the shikimate pathway, the conversion of 5-enolpyruvylshiki-
mate-3-phosphate (EPSP) to chorismate [13].
Brucella suis, a gram-negative bacteria, which is responsible for
animal brucellosis in a variety of mammalian hosts, had the aroC-
encoded CS gene disrupted in an array for oral vaccine against bru-
cellosis. The B. suis mutant is highly attenuated in tissue culture
and murine virulence models [14]. Other pathogenic bacteria as
Salmonella enterica and Yersinia enterocolitica with aroC knock-
out [15-16] also grow slowly in vivo. The in vivo behavior of the
bacterial aroC mutants (low-level residual virulence) suggests the
importance of the CS for these microorganisms and shows that
mutants can be candidate vaccines and the protein can be targeted
for drug development.
The CS enzyme has been described in bacteria, fungi, plants
and apicomplexan parasites. However, the requirement of reduced
flavin cofactor is an intriguing factor, because the CS does not do
net redox change when the substrate EPSP is converted to choris-
mate. The reaction catalyzed by CS involves an 1,4-anti-
elimination of the 3-phosphate group and of the C-(6proR) hydro-
gen from EPSP, which is unusual and unique in nature [17-18]. The
determination of CS structure from Streptococcus pneumoniae
complexed with EPSP and FMN [19] could elucidate or confirm the
role of the flavin in the chemical reaction catalyzed by CS. Kinetics
studies could show that reduced flavin play a pivotal role in the
reaction. In the structure of S. pneumoniae, EPSP molecule is
stacked above the si-face of the isoalloxazine ring in an average
distance of 3.3 Å. This close juxtaposition of FMN and EPSP sup-
ports earlier suggestions for a direct role of reduced FMN in the
elimination reaction, such as a radical mechanism, in which elec-
tron transfer from the reduced cofactor to the substrate initiates C-O
bond breakage of the phosphate group to yield a substrate-derived
neutral radical [18, 20-24]. As there is not net redox change in the
reaction, a reverse transfer of the transiently donated electron is
required to complete the catalytic cycle [18]. This unusual reaction
catalyzed by CS also makes it unique among flavin-dependent en-
zymes.
CS FROM MICROORGANISMS (BACTERIA AND FUNGI)
CS was first described from three different microbial sources:
Escherichia coli [12], Bacillus subtilis [25] and Neurospora crassa
[26] (a gram-negative bacteria, a gram-positive bacteria and a
fungi). Initially, it was believed that these identified proteins were
three different enzymes because they were biochemically different.
However, although they have many differences in their biochemical
proprieties and molecular masses, these enzymes catalyze the same
reaction, the conversion of EPSP to chorismate. Moreover, it had
been observed that there was another feature common among the
three enzymes: the reduced flavin requirement. It was observed that
the E. coli CS is active only in anaerobic conditions in the presence
of either chemically or enzymatically reduced flavin [12], while the
enzyme of N. crassa and B. subtilis appeared to be associated with
a second enzymatic activity, the NAD(P)H dependent-flavin reduc-
tase [25, 27-29]. However, these two organisms use the NAD(P)H
differently. The N. crassa enzyme possesses an intrinsic flavin re-
ductase activity, located in the same polypeptide chain of CS. On
the other hand, the B. subtilis CS appears associated to two other
enzymes: the 3-dehydroquinate synthase, the second enzyme of
shikimate pathway, and a NAD(P)H dependent-flavin reductase,
forming a heterotrimeric complex. Thus, the CS from N. crassa,
owing to this additional capacity to reduce flavin is termed bifunc-
tional CS, while the other two CS, which do not have additional
capacity to reduce flavin, are termed monofunctional CS.
Initially, it was believed that the difference between molecular
weight of monofunctional E. coli
CS enzyme (M
r
39,138 Da) and
bifunctional N. crassa CS enzyme (M
r
46,400 Da) was due to a
domain responsible for reduction of flavin [30]. Based on sequence
comparisons with monofunctional CSs, two regions of 18 internal
amino acids residues and 29 C-terminal amino acids residues
unique to N. crassa were deleted. The presence of these two regions
in CS polypeptide chain was found not to be essential by comple-
mentation with an E. coli strain lacking CS [31]. The further char-
acterization of Saccharomyces cerevisiae CS has reported a smaller
(40,800 Da) and yet bifunctional enzyme, which complies with the
hypothesis that is not feasible to predict the mono or bifunctionality
based on the molecular weight [32]. At present, all fungi CSs
known are bifunctional, while all bacterial and plant CSs are mono-
functional. However, it is not known exactly what residues or re-
gion of active site are responsible for the bifunctionality of fungi
CS, despite the experimental efforts trying to determine these prop-
erties.
CSS EVOLUTION
One question that remains without answer is how was the
mono/bifunctional CSs evolution? A possible explanation is the
existence of a common ancestor with flavin reductase activity [32].
It is has been concluded from an earlier phylogenetic analysis that
chorismate synthases are monophyletic [33], it is not know to date
whether the ancestral chorismate synthase is mono – or bifunctional
[34]. However, it has been suggested that the common ancestor was
probably bifunctional given that it is difficult to image the evolution
of the intrinsic reductase activity in a framework of monofunctional
enzymes [33,34]. It was surmised that bifunctionality may have
either been maintained only in organisms in which the availability
of reduced flavin is limiting or perhaps there was positive selection
of monofunctionality. Thermotoga maritima is thought to be one
the oldest eubacterium and appears to have undergone considerable
lateral gene transfer from the archaea. A phylogeneic tree of all CSs
presently known suggests that T. maritima CS diverged with the
archaea and moreover, considerably before any of the CSs for
which bifunctionality is known (fungi CSs). Thus the classification
of its CS as monofunctional could be considered to be cognate to
the ancestral CS and would therefore not lend support to bifunc-
tionality being ascendant [34]. Furthermore, the plant and fungal
CSs emerged from a common ancestral protein after diverging from
the monofunctional bacteria proteins. In this case, if sufficient fla-
vin is available in the cell, there may have been no selective pres-
sure to maintain de flavin reductase activity in some species. By
now, all bifunctional CSs known are from fungal origin, showing
that among these organisms the bifunctionaly persisted under selec-
tive pressure, i.e. in organisms where the availability of reduced
flavin is limiting for growth [33]. The shikimate pathway in fungi
has an interesting feature: the presence of arom complexes - a pen-
tafunctional protein complex that catalyzes the second to sixth reac-
tion of the pathway [35]. Whether there is a causal relationship in
the presence of arom and the bifunctionality of fungal CSs remain
to be solved.
Nowadays several microbial CSs were isolated from different
sources such a M. tuberculosis [36], Helicobacter pylori [37], Sal-
monella typhimurium [38], S. cerevisiae [39] and Staphylococcus
aureus [40] and these could help to solve problem of CS evolution.
APICOMPLEXAN PARASITES
The herbicide glyphosate, a potent inhibitor of the shikimate
pathway enzyme EPSP synthase, has been shown to inhibit the in
vivo growth of the apicomplexan parasites Toxoplasma gondii, P.
falciparum and Cryptosporidium parvum. This effect has provided
evidence for the presence of shikimate pathway in apicomplexan
parasites. The CS was the first shikimate pathway enzyme identi-
fied in apicomplexan parasites by anaerobic assay [9].
The genome projects have identified the ORFs from
aroC-
encoded CS gene by sequence homology in apicomplexan parasites.
Many introns were found in aroC gene from T. gondii, but none
was identified in P. falciparum aroC gene. The lack of introns al-
lows the direct expression of P. falciparum CS in bacterial systems.
129
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Chorismate Synthase: An Attractive Target For Drug Development Current Drug Targets, 2007, Vol. 8, No. 3 3
Moreover, the T. gondii aroC is in a single copy, which allows gene
deletion experiments. Other interesting feature in the T. gondii aroC
gene is a potential GCN4 binding motif (TGACTC) in the 5' non-
coding region [3]. This specific upstream promoter element is pre-
sent in genes that are regulated by the control-activator protein
GCN4. In S. cerevisiae the expression of genes encoding shikimate
pathway enzymes DAHP synthase (3-deoxy-D-arabino-heptulo-
sonate 7-phosphate synthase) and the pentafunctional arom com-
plexes (which catalyzes the steps second to sixth in the pathway)
are regulated by GCN4 protein [41], which indicates a possible
regulation of shikimate pathway by CS expression in apicomplexan
parasites.
In plants, the shikimate pathway enzymes, like other nuclear-
encoded plastid enzymes, are post-translationally targeted to the
plastid by an amino-terminal leader sequence. The apicomplexan
parasites have a plastid-like organelle called apicoplast [42]. Al-
though the apicoplast proteins from apicomplexan parasites also
have leader sequences, there is no evidence that shikimate pathway
enzymes are localized in the apicoplast. Neither the T. gondii CS
gene nor P. falciparium CS gene have an obvious amino-terminal
leader sequence similar to that found on CSs from higher plants,
which strongly suggests that CS apicomplexan enzymes are in the
cytosol, and not in the apicoplast [43]. The shikimate pathway in
fungi also occurs in cytosol and phylogenetic analysis has demon-
strated that CSs from apicomplexan parasites are closer to fungal
CSs than plants CSs, which is consistent with the idea of a non-
apicoplast CS. Moreover, it also indicates a common ancestor to
fungal and apicomplexan enzymes [44].
Studies with CS from P. falciparium have characterized the en-
zyme biochemically and immunologically. Cofocal immunofluo-
rescence and cellular fractions followed by Western blot analysis
has shown that CS is located in the cytosol of this organism in dif-
ferent stages of infection, which is in agreement with the phyloge-
netic analysis. However, despite the close relationship with fungal
proteins, the enzymatic assay has demonstrated that P. falciparum
CS is monofunctional as in plants and bacteria [45]. Therefore, the
CS from P. falciparum has a combination of properties from plant
and fungal enzymes and it seems to possess properties distinct from
previously described CSs. It was found that both T. gondii and P.
falciparum enzymes differ from other known CS in possessing a
number of unique insertions [3]. Although the CSs in apicomplexan
parasites are larger in length, they share all of the amino acids to be
highly conserved in CSs from other species [9]. The implication of
a cytosolic location for the development of new drugs for apicom-
plexan parasites therapy is clear, fewer membranes will have to be
crossed.
The inhibition of P. falciparum growth with a dsRNA encoding
a 900 bp fragment of aroC has shown that CS is required for the
normal growth of this pathogen [46]. It suggests that CS may be a
viable target for chemotherapy. Moreover, two fluorinated analogs
of shikimate: (6R)-6-fluoroshikimate and (6S)-6-fluoroshikimate
have been shown to be inhibitors of the P. falciparum growth in
vitro [47]. The possible function of apicomplexan CS in shikimate
pathway regulation makes this enzyme a promising target for drug
development. Increasing our understanding of the apicomplexan
shikimate pathway enzymes may expedite the discovery and design
of new antiparasitic agents.
CS MECHANISM
The CS-catalyzed chemical reaction was first described in E.
coli extracts. It was reported that CS was inactivated under aerobic
conditions and it could be activated most effectively by a reduced
flavin cofactor [12]. Although many flavoproteins are involved in
redox reactions, there is not an overall redox when EPSP is con-
verted to chorismate and the reduced flavin mononucleotide is not
consumed during the reaction. Another interesting feature in CS
catalysis is an unusual 1,4 anti elimination of 3-phosphate group
and the C(6proR) hydrogen from EPSP, implied in a nonconcerted
reaction, which is interesting because it involved the cleavage of a
nonactivated C-H bond [17]. The strict requirement for a reduced
flavin mononucleotide and the anti 1,4 elimination are both unusual
aspects which make CS mechanism unique among flavin-dependent
enzymes [33]. Although there are two different classes of CS, dis-
tinguished by their ability to use NADPH for the reduction of FMN,
there is no apparent difference in the mechanism of elimination
reaction [18].
A nonconcerted mechanism was demonstrated by studies that
include a secondary tritium kinetic isotopic effect at C(3) [48], tran-
sient kinetic studies [20] and a secondary deuterium kinetic iso-
topic effect at C(4) [49]. The kinetic isotopic experiments have
revealed that C(6)-H and C(3)-O bonds are cleaved in distinct steps,
and the phosphate group is eliminated before the C-H cleavage.
Two intermediates of nonconcerted mechanism have been proposed
to CS reaction: the cationic and the radical intermediate [50]. Al-
though reported isotopic effects could not clearly distinguish be-
tween the radical and cationic mechanisms, they have provided
evidence against alternative mechanisms. Moreover, the observa-
tion of spectral changes associated with the flavin during the cata-
lytic reaction has provided the strong evidence that the reduced
flavin has been actively involved in the reaction mechanism [51].
The role of flavin mononucleotide in catalysis derives from the lack
of activity of the bifunctional enzyme from N. crassa with 5-deaza-
FMN [52] and the formation of a stable flavin semiquinone radical
with the monofunctional E. coli enzyme and the substrate analog
(6R)-6-fluoro-EPSP, as a ternary complex [23, 52].
The characterization of the flavin intermediate has shown that it
is not simply associated with the conversion of substrate to product
since it must be formed before any chemical step involving EPSP.
The formation of a flavin intermediate is associated with the forma-
tion of the ternary complex between enzyme, EPSP and reduced
flavin, as described above. The flavin intermediate is generated
very rapidly within a few milliseconds and disappears when all
substrate has been consumed [20]. Spectral observations have
shown that the reduced flavin is bound to CS in its monoanionic
(deprotonated) form in the absence of substrate [22]. On the other
hand, studies with flavin analogs have demonstrated that CS binds
preferably the flavin in a neutral form, including the protonated
reduced form, rather than anionic forms in the presence of EPSP or
(6R)-6-F-EPSP [21]. The reported data led to the conclusion that
CS binds flavin in its monoanionic form and the protonation of the
N(1) position of FMN occurs upon binding of either EPSP or (6R)-
6-F-EPSP to the binary complex, which indicates that protonated,
reduced flavin is the observed reaction intermediate. The protona-
tion could lower the reduced flavin redox potential, converting the
reduced flavin into a better reductant, as required to promote the
radical chemistry [20]. Moreover, chorismate was produced from
(6R)-6-F-EPSP with the formation of a flavin semiquinone radical,
indicating the involvement of radical chemistry in this reaction. The
ternary complex is formed and a single electron is transferred from
the low potential flavin to the analogue, which promotes the elimi-
nation of a phosphate group to generate the allylic radical [23]. This
explanation is in agreement with the essential requirement of a fla-
vin co-factor.
Studies with substrate analogs have shown that even though the
(6S)-6-F-EPSP is a competitive inhibitor in N. crassa CS [54], it is
converted to 6-fluorochorismate in E. coli CS at a rate between 270
and 370 times slower than EPSP [50]. Studies with fluorinated sub-
strate analogues have shown that in E. coli CS the C(6)-H bond
breaking occurs after the formation of the flavin intermediate. Al-
though the C(6)-H breaking does not contribute significantly to rate
limitation in E. coli enzyme, it is partially rate-limiting with N.
crassa enzyme [52 54]. On the other hand, this result and deuterium
kinetic isotopic effect studies [55] have provided evidence that
steps after flavin intermediate decay, which may include product
130
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
4 Current Drug Targets, 2007, Vol. 8, No. 3 Dias et al.
release, are not significantly rate-limiting with EPSP or (6S)-6-F-
EPSP.
Spectroscopic and physical studies have shown that E. coli CS
undergoes a major structural change when the ternary complex is
formed. When the EPSP and FMN are bound, the enzyme has a
more compact shape, designated as closed form. On the other hand,
the apoenzyme probably exhibits more conformational flexibility,
designated as opened form [56]. The elucidation of S. pneumoniae
CS structure provides evidence that the binding of EPSP induces
changes in protein conformation, concurrent with a protonation of
FMN to give the neutral reduced form [19]. It is observed mainly
by the clear change in the position of His
110
when the opened and
closed structures are compared. It was proposed that the His
110
was
the probable residue that protonates the FMN and neutralizes the
reduced FMN in the first step of the reaction. The His
10
also ap-
pears to play a role in the reaction by providing a proton to neutral-
ize the charge on the departing phosphate group [19].
The fundamental role of these two histidine residues in CS ac-
tive site, as observed in S. pneumoniae, was confirmed by site-
directed mutagenesis of two histidine residues in the active site of
N. crassa CS. These histidines, His
106
and His
17
, were replaced by
alanine, reducing the activity 10- and 20-fold, respectively. The
phosphate release occurs before of the C-(6proR) hydrogen cleav-
age [20] and probably the two histidines residues are involved in
chemical steps leading directly to the initial C-O bond cleavage. It
was proposed that His
106
is involved in the protonation of monoan-
ionic reduced FMN, so it functions as the acid that donates the pro-
ton to N(1). This protonation may serve to direct the electron den-
sity so that charge transfer to the double bond of the substrate is
maximized. The function of His
17
is protonation of the leaving
phosphate group of EPSP. A detailed mechanism for CS reaction
was proposed based on structure and site-directed mutagenesis.
According to this model the EPSP binds to enzyme-bound reduced
FMN, forming a closed form. Then, an electron is transferred to the
EPSP double bond, initiating C-O bond cleavage and the release of
phosphate. In this step the His
17
acts as a general acid to neutralize
the incipient charge on the oxygen atom. The resulting C(4a)-
neutral flavin semiquinone tautomerizes to a radical species where
the unpaired electron residues on N(5) with concomitant abstraction
of the (proR-H) from C(6) by the N(5) [18]. Following the cleavage
of phosphate and possibly partial re-opening of the active site, the
intermediate remains bound while the proton is abstracted. The
atom N(5) of FMN is positioned directly below the C(6-proR) hy-
drogen of EPSP and is ideally positioned to remove this proton
from the transient intermediate. In the last step, the deprotonation of
the reduced flavin restores the initial state of the cofactor. The re-
generation of reduced FMN occurs with the reopening of the active
site, involving deprotonation by His
110
in the S. pneumoniae [19]. It
is also proposed that neither of the two histidines residues is essen-
tial for binding of the oxidized cofactor, representing only a role in
the stabilization of the change of the cofactor [18].
The future contributions as studies with substrate analogues,
high-resolution structures information and site-directed mutagenesis
in other invariant amino acids residues in the active site will pro-
vide further insights into this unprecedented flavin-dependent reac-
tion mechanism.
THE ROLE OF NADPH IN THE BIFUNCTIONAL CHO-
RISMATE SYNTHASE
CSs from fungal sources, i.e. S. cerevisiae and N. crassa, have
and additional intrinsic catalytic activity, which utilizes NAD(P)H
to reduce the flavin cofactor. (13, 56). The NAD(P)H is utilized in
the bifunctional CS to generate the essential reduced FMN cofactor.
NAD(P)H is consumed for the reduction of FMN followed by re-
oxidation of the reduced FMN by molecular oxygen. The rate of
flavin reduction probably is independent of the concentration and
presence of EPSP. The most intriguing question about the bifunc-
tional CS is the nature and the location of the NAD(P)H binding
site. All studies carried out to identify the sequence responsible for
binding of NAD(P)H have failed (31), therefore, the ability of bi-
functional CS to bind NAD(P)H is embedded in the protein struc-
ture revealing an unusual mode of interaction between these CS and
NAD(P)H [13]. Furthermore, Kitzing et al. (2001) [13] have ob-
served that EPSP and NAD(P)H compete for a common binding
site because the rate of NAD(P)H oxidation strongly depends on the
concentration of EPSP. It is conceivable because both NAD(P)H
and EPSP possess a phosphate group, which in turn plays an impor-
tant role in binding to the active site of CS. The site-directed
mutagenesis studies of important histidine residues of the active site
of N. crassa CS demonstrated that specific histidines to alanines
replacements affected the intrinsic NAD(P)H:FMN oxidoreductase
activity. It is mentioned that the lack of activity of the two histidine
to alanine mutant proteins under aerobic assay conditions does not
lead to conversion of EPSP to chorismate. Although both amino
acid exchanges affect the intrinsic oxidoreductase activity of CS,
they are apparently not decisive for secondary enzymatic activity,
as they do not abolish the flavin reductase activity [18]. Thus, the
amino acid residues that bring about bifunctionality remain obscure
[18]. Therefore, only the determination of the three-dimensional
structure of a bifunctional CS complexed with NAD(P)H will pro-
vide a description at molecular level of the essential residues re-
sponsible for bifunctionality.
THE STRUCTURE OF CHORISMATE SYNTHASE
Macheroux et al. (1998) [56] have predicted the CS structure in
a - barrel form as the result of secondary structure prediction
efforts. However, the determination of high-resolution structures by
crystallography has shown that they were not correct and the CS
structure is very different from what was expected. The small-angle
X-ray scattering data have indicated that the enzyme presents a
more compact overall shape in the ternary complex when both oxi-
dized FMN and EPSP are bound to the enzyme as compared to the
apoenzyme. Probably, the presented apoenzyme exhibits more con-
formational flexibility, even though these ligands have no effect on
the oligomerization state of the enzyme [5].
All crystal structures of CS solved to date show that the quater-
nary structure is tetrameric with approximately 222 symmetry (Fig.
(1a)). Moreover all structures have high structural similarity repre-
senting other distinct class of a conserved enzyme family, since it
lacks a primary sequence similarity to other classes of enzymes.
Despite the CSs being tretameric in its crystal form, in solution it
can be present as a dimer, tetramer or dimer/tetramer equilibrium
[12, 30, 39-40, 58].
The organisms whose CS crystal structures have been deter-
mined are: S. pneumoniae [19], Aquifex aeolicus [59], H. pylori
[60], S. cerevisaie [61], Campylobacter Jejuni [PDB access code:
1SQ1]. Furthermore preliminary data of X-ray crystallography for
CS from M. tuberculosis have also been reported [36].
The dominant structural topology
of CS monomer is character-
ized by a -- sandwich, in which each monomer consists of a
central helical core formed for four long -helices, sandwiched
between two four-stranded antiparallel beta sheets which is sur-
rounded by loops and discrete stretches of helix and sheet .
These layers are packed to form a compact structure [19, 59-61]
(Fig. (1b)). The structure could also be described as two intimately
contacted sub-domains, consisting of anti-parallel sheet covered by
helices formed from -2 halves that are related by pseudo 2-fold
symmetry. However, the lack of identity between the sequences,
which form two domains, shows that could not have occurred for
gene duplication events [61-62].
131
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Chorismate Synthase: An Attractive Target For Drug Development Current Drug Targets, 2007, Vol. 8, No. 3 5
Although CS is a tetramer in the crystals, it is better character-
ized as a dimer of dimers in its basic structural unit because two
dimers compose the observed tetramer [19]. The dimer interface is
extensive and includes antiparallel interactions between the ends of
the sheet 2 from each subunit to give an extended eight-stranded
sheet spanning both subunits (Fig. (1c)) [62].
ACTIVE SITE CHARACTERISTICS
The interesting aspect of active site is the large number of loop
regions that comprise the substrate binding pocket (fig. (1d)). Two
of them are provided by the adjacent subunit across the dimer inter-
face. In the ternary complex, FMN is almost completely buried,
with part of its accessible surface area being occluded by the sub-
strate, EPSP. Thus, the structure indicates an ordered binding of
FMN followed by EPSP, which is consistent with previous rapid
kinetic studies [56].
The FMN-binding site in CS is unique; no other flavoenzyme
presents such characteristic. In the structure of S. pneumoniae, the
FMN molecule makes few specific polar interactions with protein,
most part of them are contacts between hydroxyl and phosphate
oxygens of the ribityl chain. The flexible regions that surround the
FMN site are rich in strictly or highly conserved residues. The
FMN phosphate sits at the dimerization interface and it makes a
number of contacts with residues from the adjacent monomer. Fur-
thermore, there are a considerable number of solvent molecules
close to both the phosphate and ribityl regions of FMN. The water
molecules are discrete and ordered. They are responsible for medi-
ate interactions between FMN and the surrounding residues. The
isoalloxazine ring system of FMN is hydrophobic and makes few
specific interactions with the protein. However, it buries a consider-
able area of hydrophobic surface by packing the re-face against the
complementary surface of the protein [19]. This environment con-
tributes to the remarkable lowering of the flavin´s reduction poten-
tial to a value comparable to the most reducing flavodoxins [62]. A
notable feature of the bound FMN is a significant deviation from
planarity displayed by isoalloxazine ring. The pyrimidine ring and
the dimethylbenzene ring make an angle of approximately 10º and
the re-face of the isolloxazine ring is convex. It appears to assist the
substrate EPSP-binding site. This characteristic is observed in very
few other flavoenzymes [60, 63-65]. Furthermore the FMN binding
site has a number of positive charges in the proximity of the isol-
loxazine ring, which can increase the redox potential of FMN,
whereas a negative charge or a hydrophobic environment is ex-
pected to lower it. The positive charges at this location can possibly
stabilize negatively charged N1 in the reduced flavin, thus keeping
the electron until the substrate EPSP is bound [60].
The EPSP molecule is oriented relative to the N (5) position of
the isoalloxazine ring for electron transfer and hydrogen atom ab-
straction to and from the substrate, respectively. The EPSP mole-
cule is attacked above the si-face of the isoalloxazine ring in an
average distance of 3.3 Å as in the structure of CS from S. pneuno-
miae (Fig. (1d)). Owing to this fact, it can be observed that during
the reaction there is no accumulation of flavin radical intermediates
on the stopped-flow millisecond time scale. Hence, CS catalyzes a
Fig (1). Structure of CS; A)Tetramer structure; B) Monomer structure; C) Dimer; D) structure of active site.
132
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
6 Current Drug Targets, 2007, Vol. 8, No. 3 Dias et al.
unique reaction when occurs only the transfer of one electron,
which is unusual among flavoenzymes. The localization of the
FMN and EPSP in the CS has provided evidence that the flavin is
well placed to abstract a hydrogen atom from the 6-proR position of
the EPSP molecule, rather than just an electron with the proton
being accepted by an amino acid side chain. This form of flavin can
donate an electron and subsequently accept a hydrogen atom, stabi-
lizing again [19, 62].
In the structure of CS from S. pneumoniae, the EPSP molecule
makes a number of polar interactions with the protein, and few
hydrophobic contacts. The site of EPSP is hydrophilic and has an
environment extensively basic with many basic residues, mainly
arginines around of the binding site, which are extremely conserved
in many sequences, thereby showing its importance in the charge
stabilization of EPSP molecule [19].
Maclean and Ali (2003) [19] observed that the quaternary struc-
ture of S. pneumoniae has two conformational states. One of the
four monomers of the tetramer has differences when compared with
the other three. These changes occur principally in two loops of
active site comprised of highly conserved residues. The quaternary
structure shows one monomer with the active site more accessible
having therefore an open conformation, whereas the other three
have closed conformations. These differences provide vital infor-
mation about the mobility of these active site loops, which play an
important mechanistic role [19]. These loops have important basic
residues related with the catalytic mechanism of the protein. Differ-
ences between open and closed active sites demonstrate that the
side chain of one histidine located in one of these loops has some
conformational flexibility, and interacts with the FMN in the open
form but with O12 of EPSP in the closed form. Furthermore other
highly conserved residues in these loops also have different con-
formations in the two states showing that these residues play impor-
tant role in the catalytic mechanism [19]. Ahn et al. (2004) [60]
have observed also that the structure from H. pylori is surrounded
by flexible regions that are also highly conserved. Ahn et al. (2004)
[60] crystallized the apoenzyme and CS complexed with FMN. The
authors have observed that FMN could cause conformational
changes in these loops in CS from H. pylori, furthermore, they have
observed that the movement of these regions results in a more apo-
lar environment for the bound cofactor [60]. This structural infor-
mation confirms the hypothesis, in which the CS undergoes a major
change on binding oxidized flavin and EPSP and that the apoen-
zyme exhibits more conformational flexibility than ternary com-
plex. Furthermore it was observed that alterations occurred in the
visible CD spectrum of the enzyme-bound FMN on binding of
EPSP showing that the environment of the flavin changes consid-
erably [56].
STRATEGY FOR DEVELOPMENT DE DRUG BASED IN
THE CHORISMATE SYNTHASE
Based on the structural knowledge of the enzymes of the shiki-
mate pathway and their mechanisms of catalysis, it is possible to
design potential inhibitors in a rational way against pathogenic
bacteria, fungi and apicomplexan parasites. For amenable manner
and for knowledge about the shikimate, it was chosen for evaluating
of the design potential inhibitors. The strategy used was to incorpo-
rate fluorine at regio- and stereo-specific site of this molecule [66-
68]. With the knowledge of loss of C(6-proR) hydrogen in the reac-
tion catalyzed for CS [17, 66], two shikimate analogs were synthe-
sized, the (6R)-6-fluoro- and (6S)-6-fluoro shikimate acids. The
fluorine in (6R)-6-fluoro shikimate acid occupies the position of the
hydrogen abstracted during conversion of EPSP to chorismate. The
(6R)-6-fluro- and (6S)-6-fluro derivatives from shikimate acid are
metabolized by shikimate kinase and then to EPSP synthase to give
the corresponding diastereoisomeric 6-fluoro-EPSP. The (6R)-6-
EPSP might inhibit the CS, while the (6S)-6-EPSP would be proc-
essed further and might lead to lethal synthesis [66]. The (6R)-6-
fluoro-EPSP binds to E. coli CS in vitro and causes the oxidation of
enzyme-bound flavin, but it is not converted further, and neither
chorismate nor any other product could be detected [53]. On the
other hand, the (6S)-6-fluoro-EPSP is slowly converted to choris-
mate by CS from E. coli to a product identified as 6-fluoro-
chorismate [66] that inhibits the 4-aminobenzoic acid synthesis [55,
66]. Because of its greater potency in vitro, the shikimic acid ana-
logues were tested for antibacterial activity in vivo. The (6S)-6-
fluoro-shikimate was more protective against bacterial intraperito-
neal challenges in a mouse protection test. Mice were infected with
S. aureus and Salmonella dublin, but none of these strains would
grow when mice were treated with (6S)-6-fluro-shikimate on the
minimal chemically defined medium [66, 69].
The 6R- and 6S- isomers of 6-fluoro-shikimates have antibacte-
rial activity displaying minimum inhibitory concentrations of 64
and 0.5 μg mL
-1
against E. coli, respectively [66]. Owing to the lack
of (6proR) hydrogen in (6R)-6-fluoro-EPSP, it cannot be considered
a substrate of CS. Thus, this compound promotes one electron oxi-
dation of the E. coli enzyme [53] and it is a competitive inhibitor of
the N. crassa enzyme. The (6S)-6-fluoro-shikimate is also a com-
petitive inhibitor of the N. crassa enzyme, because there was no
conversion of this analog to 6-fluorochorismate by the enzyme of
this organism [54]. On the other hand, in CS from E. coli, this ana-
logue is not an inhibitor, since the conversion to 6-fluorochorismate
occurs. It can occur due the N. crassa enzyme to be different, once,
it is bifunctional [30, 57], differently of the E. coli enzyme. Despite
of conversion of (6S)-6-fluoro-EPSP for E. coli CS, it is converted
in a rate between 270 and 370 times slower than EPSP [50].
The decreased rate of reaction is consistent with the electron
withdrawing fluoro constituent destabilizing an allylic cationic
intermediate [70] that would be generated by the loss of phosphate
from the substrate [50]. There is a strong evidence for a stepwise,
rather than a concerted condensation reaction, involving an allylic
cationic intermediate with the fluoro substituent adjacent to the
allylic system. An allylic cationic intermediate in the CS reaction
would also be destabilized, in part, by negative hyperconjugative
effect of the adjacent fluoro substituent [71]. The negative hyper-
conjugative effect of the fluorine could destabilize an allylic radical
intermediate that would be formed by a mechanism involving an
additional one-electron reduction [50]. Once a ternary complex is
formed between the reduced holoenzyme and the (6R)-fluoro-EPSP
analogue, a single electron is transferred from the low potential
flavin to the analogue. This promotes the elimination of the phos-
phate group to generate an allylic radical. This intermediate is inca-
pable of transferring an electron and proton to the flavin from the
6R position due to its (6R)-fluoro group [23]. The (6S)-6-fluoro-
shikimate does not inhibit the CS, but the product formed by CS,
the 6-fluorochorismate inhibits the PabA and PabB. These enzymes
are responsible for conversion of chorismate to 4-aminobenzoate
precursor 4-amino-4-deoxychorismate [72].
The analogues of flavin also could be used to the design of in-
hibitors of CS, since it has been demonstrated that occurs lack of
activity of the enzyme with reduced 5-deaza-FMN [23].
Other class of inhibitors of CS was developed by Thomas et al.
(2003) [73]. Among these inhibitors are the benzofuran-3[2H]-one.
This inhibitor is moderately potent, with IC
50
of 8 μM. Kinetic
studies demonstrated that the benzofuran-3-[2H]-one is a competi-
tive inhibitor with respect to EPSP and noncompetitive with FMN,
suggesting that the compound binds in the active site of the en-
zyme. However, owing to the metabolic vulnerability and potential
toxicity of the 6,7-dihydroxy functional group, and also in an effort
to increase potency, further synthetic efforts have been undertaken
[73].
CONCLUSION
CS is an important enzyme of the shikimate pathway, having a
critical role in metabolism of microorganisms. Furthermore it is
133
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
Chorismate Synthase: An Attractive Target For Drug Development Current Drug Targets, 2007, Vol. 8, No. 3 7
fascinating, mainly for their unusual catalytic mechanism and for
their unique folding in nature. The knowledge about the CSs has
been evaluated day by day. The determination of various structures
of this protein would contribute to an understanding of the details of
the mechanism of catalysis and reveal the role played by important
active site residues, such as the His
10
and His
110
in S. pneumonie.
Future research could also contribute to the understanding of the
unprecedented flavin-dependence in the reaction mechanism, which
can be through of site-directed mutagenesis and the utilization of
substrate and cofactor analogues. The major challenge today, is to
precisely determine the role of NAD(P)H in bifunctional CSs and
the residues that are involved in the their binding. On the other
hand, today, major new opportunities in structure-based drug design
are possible due the determination of CS structure complexed with
both substrate and cofactor. This knowledge could contribute to the
development of new drugs including the use of combinatorial
chemistry and virtual screening methods that could render inhibi-
tors more specific to compete with substrate and cofactor in binding
to the enzyme. The development of new drugs or medicines against
infectious disease such as tuberculosis, respiratory disease and ma-
laria will help hundreds of millions of people worldwide [74-86],
mainly in developing countries, where poor people will most bene-
fit from these efforts.
ACKNOWLEDGMENTS
This work was supported by grants from FAPESP
(SMOLBNet, Proc. 01/07532-0, 03/12472-2, 04/00217-0) and Mil-
lenium Institute, MCT/CNPq to DSS and LAB. MSP, DSS, WFA
and LAB are Research Awardees from the National Research
Council.
ABBREVIATION LIST
CS = Chorismate synthase
TB = Tuberculosis
EPSP = 5-enolpyruvylshikimate-3-phosphate
FMN = Flavin mononucleotide
FMN
red
= Flavin mononucleotide reduced
FMN
oxi
= Flavin mononucleotide oxidized
FMN
SQ
=
Flavin mononucleotide semiquinone
DAHP = 3-deoxy-D-arabino-heptulosonate 7-
phosphate
6(R)-6-F-EPSP = (6R)-6-fluoro-5-nolpyruvylshikimate-3-
phosphate
6(S)-6-F_EPSP = (6S)-6-fluoro-5-nolpyruvylshikimate-3-
phosphate
His
10
= Histidine 10
His
110
= Histidine 110
PDB = Protein data bank
REFERENCES
[1] Trouiller, P.; Torreele, E.; Olliaro, P.; White, N.; Foster, S.; Wirth,
D. and Pécoul, B. (2001) Trop. Med. Internat. Health., 6(11), 945-
951.
[2] Corbett, E. L.; Watt, C. J.; Walker, N.; Maher, D.; Williams, B. G.;
Raviglione, M. C. and Dye, C. (2003) Arch. Intern. Med., 1639,
1009-1021.
[3] Roberts, C. W.; Roberts, F.; Lyons, R. E.; Kirisits, M. J.; Mui, E.
J.; Finnerty, J.; Johnson, J. J.; Ferguson, D. J. P.; Coggins, J. R.;
Krell, T.; Coombs, G. H.; Milhous, W. K.; Kyle, D. E.; Tzipori, S.;
Barnwell, J.; Dame, J. B.; Carlton, J. and McLeond, R (2002) J. In-
fect. Dis., 185, S25-36.
[4] coul, B.; Chirac, P.; Trouiller, P. and Pinel, J. (1999) J. of the
American Med. Association, 281, 361-367.
[5] Coates, A.; Hu, Y.; Bax, R. and Page, C. (2002) Nat. Rev. Drug.
Discov., 1, 895 – 910.
[6] Bentley, R. (1990) Crit. Rev. Biochem. Mol. Biol., 1990, 307-384.
[7] Herrmann, K. M. (1995) Plant Cell, 7, 907-919.
[8] Parish, T. and Stoker, N. G. (2002) Microbiol., 148, 3029-3077.
[9] Roberts, F.; Roberts, C. W.; Johnson, J. J.; Kyle, D. E.; Krell, T.;
Coggins, J. R.; Coombs, G. H.; Milhous, W. K.; Tzipori, S.;
Ferguson, D. J. P.; Chakrabarti, D. and McLeond, R. (1998) Na-
ture, 393, 801-805.
[10] Gibson, M.I. and Gibson, F. (1962) Biochim. Biophis. Acta, 65,
160-163.
[11] Ratledge, C. (1982) Nutrition, growth and metabolism in: Ratledge,
C., Stanford, J. L. (Eds.), The biology of the Mycobacteria, vol. 1,
Academic Press, London, pp. 185-271.
[12] Morell, H.; Clark, M.J.; Knowles, P.F. and Sprinson, D.B. (1967) J.
Biol. Chem., 242, 82-90.
[13] Kitzing. K.; Macheroux, P. and Amrhein, N. (2001) J. Biol. Chem.,
276(46), 42658-42666.
[14] Foulongne, V.; Walravens, K.; Bourg, G.; Boschiroli, M.L.; God-
froid, J.; Ramuz, M. and O'Callaghan D. (2001) Infect. Immun.,
69(1), 547-550.
[15] Bowe, F.; O'Gaora, P.; Maskell, D.; Cafferkey, M. and Dougan, G.
(1989), Infect. Immun., 57, 3234-3236.
[16] Lowe, D.C.; Savidge, T.C.; Pickard, D.; Eckmann, L.; Kagnoff,
M.F.; Dougan, G. and Chatfield, S.T. (1999) Infect. Immun., 67,
700-707.
[17] Floss, H. G.; Onderka, D. K. and Carroll, M. (1972) J. Biol. Chem.,
247, 736-744.
[18] Kitzing, K.; Auweter, S.; Amrhein, N. and Macheroux, P. (2004) J.
Biol. Chem., 279, 9451-9461.
[19] Maclean, J. and Ali, S. (2003) Structure, 11(12), 1499-1511.
[20] Bornemann, S.; Lowe, D. J. and Thorneley, R. N. (1996) Biochem-
istry, 35(30), 9907-9916.
[21] Macheroux, P., Bornemann, S.; Ghisla, S. and Thorneley, R. N.
(1996) J. Biol. Chem., 271(42), 25850-25858.
[22] Macheroux, P.; Petersen, J.; Bornemann, S.; Lowe, D. J. and Thor-
neley, R. N. (1996) Biochemistry, 35(5), 1643-1652.
[23] Osborne, A.; Thorneley, R. N.; Abell, C. and Bornemann, S. (2000)
J. Biol. Chem., 275(46), 35825-35830
[24] Machoroux, P.; Bornemann. and Thorneley, R. N. F. (1997) in
Flavins and flavoproteins 1996 (Stevenson, K. J.; Massey, V. and
Williams, C. H. J., eds) 0University of Calgary Press, Calgari,
Canada, pp. 113-122.
[25] Hasan, N. and Nester, E. W. (1978) J. Biol. Chem., 253(14), 4987-
4982.
[26] Gaertner, F. H. and Cole, K. W. (1973) J. Biol. Chem., 248(13),
4602-4609.
[27] Gaertner, F. H. (1987) Methods Enzymol., 142, 362-366.
[28] Hasan, N. and Nester, E. W. (1978) J. Biol. Chem., 253(14), 4993-
4998.
[29] Hasan, N. and Nester, E. W. (1978) J. Biol. Chem., 253(14), 4999-
5004.
[30] White P. J.; Millar, G. and Coggins, J. R. (1988) Biochem. J.,
251(2), 313-322.
[31] Henstrand, J. M.; Amrhein, N. and Schmid, J. (1995) J. Biol.
Chem., 270(35), 20447-20452.
[32] Henstrand, J. M.; Schaller, A.; Braun, M.; Amrhein, N. and
Schmid, J. (1996) Mol. Microbiol., 22(5), 859-866.
[33] Macheroux, P.; Schmid, J.; Amrhein, N. and Schaller, A. (1999)
Planta, 207, 325-334.
[34] Fitzpatrick, T. B.; Killer, P.; Thomas, R.M.; Jelesarov, I.; Amrhein,
N. and Macheroux P. (2001) J. Biol. Chem., 276(21), 18052-18059.
[35] Hawkins, A. R. and Smith, M. (1991) Eur. J. Biochem., 196, 717-
724.
[36] Dias, M. V., Borges, J. C., Ely, F., Pereira, J. H., Canduri, F., Ra-
mos, C. H., Frazzon, J., Palma, M. S., Basso, L. A., Santos, D. S.,
de Azevedo, W. F. Jr. (2006) J. Struct. Biol. 154(2) ,130-143.
[37] Ahn, H. J.; Yang J. K.; Lee, B. I.; Yoon, H. J.; Kim, H. W. and
Suh, S. W. (2003) Acta Crystallogr. D Biol. Crystallog., 59, 569-
571.
[38] Charles, I. G.; Lamb, H. K.; Pickard, D.; Dougan, G. and Hawkins,
A. R. (1990) J. Gen. Microbiol., 136(2), 353-358.
[39] Jones, D. G.; Reusser, U. and Braus, G. H. (1991) Mol. Microbiol.,
5(9), 2143-2152.
[40] Horsburgh, M. J.; Foster, T. J.; Barth, P. T. and Coggins, J. R.
(1996) Microbiology, 142, 2943-2950.
[41] Arndt, K. and Fink, G. R. (1986) Proc. Natl. Acad. Sci. USA, 83,
8516-8620.
134
Estudo Estrutural de Proteínas Alvo de Mycobacterium tuberculosis
8 Current Drug Targets, 2007, Vol. 8, No. 3 Dias et al.
[42] Köhler, S.; Delwiche, C. F.; Denny, P. W.; Tilney, L. G.; Webster,
P. and Wilson, R. J. (1997) Science, 275, 1485-1489.
[43] Mcfadden, G. I.; Reith, M. E.; Munholland, J. and Lang-Unnasch,
N. (1996) Nature, 381, 482.
[44] Keeling, P. J.; Jeffrey, D. P.; Donald, R. G. K.; Roos, D. S.; Waller,
R. F. and McFadden, G. I. (1999) Nature, 397, 219-220.
[45] Fitzpatrick, T.; Ricken, S.; Lanzer, M.; Amrhein, N. Macheroux, P.
and Kappes, B. (2001) Mol. Biol., 40(1), 65-75.
[46] McRobert, L. and McConkey, G. A. (2002) Mol. Biochem. Parasi-
tol., 119, 273-278.
[47] McConkey, G. A. (1999) Antimicrob. Agents Chemotherar., 43,
175-177.
[48] Balasubramanian, S.; Coggins, J. R. and Abell, C. (1995) Biochem-
istry, 34(1), 341-348.
[49] Bornemann, S.; Theoclitou, M. E.; Brune, M.; Webb, M. R.; Thor-
neley, R. N. and Abell, C. (2000) Bioorg. Chem., 28, 191-204.
[50] Bornemann, S.; Ramjee, M. K.; Balasubramanian, S.; Abell, C.;
Coggins, J. R.; Lowe, D. J. and Thorneley, R. N. F. (1995) J. Biol.
Chem., 270, 22811-22815.
[51] Ramjee, M. K.; Coggins, J. R.; Hawkes, T. R.; Lowe, D. J. and
Thorneley, R. N. F. (1991) J. Am. Chem. Soc., 113, 8566-8567.
[52] Lauhon, C. T. and Bartlett, P. A. (1994) Biochemistry, 33, 14100-
14108.
[53] Ramjee M. N.; Balasubramanian, S.; Abell, C.; Coggins, J. R.;
Davies, G. M.; Hawkes, T. R.; Lowe, D. J. and Thorneley, R. N. F.
(1992) J. Am. Chem. Soc., 114, 3151-3153.
[54] Balasubramanian, S. and Abell, C. (1991) Tetrahedron Lett., 32,
963-966.
[55] Bornemann, S.; Balasubramanian, S.; Coggins, J. R.; Abell, C.;
Lowe, D. J. and Thorneley, R. N. (1995) Biochem. J., 305, 707-
710.
[56] Macheroux P.; Schonbrunn, E.; Svergun, D. I.; Volkov, V. V.;
Koch, M. H.; Bornemann, S. and Thorneley, R. N. (1998) Biochem.
J., 335, 319-327.
[57] Welch, G. R.; Cole, K. W.; and Gaertner, F. H. (1974) Arch. Bio-
chem. Biophys., 165, 505-518.
[58] Schaller, A.; Windhofer, V. and Amrhein, N. (1990) Pers. Arch.
Biochem. Biophys., 282, 437-442.
[59] Viola, C.M.; Saridakis, V. and Christendat, D. (2004) Proteins,
54(1), 166-169.
[60] Ahn, H.J.; Yoon, H.J.; Lee, B. 2nd and Suh, S.W. (2004) J. Mol.
Biol., 336(4), 903-915.
[61] Quevillon-Cheruel, S.; Leulliot, N.; Meyer, P.; Graille, M.; Bre-
mang, M.; Blondeau, K.; Sorel, I.; Poupon, ª; Janin, J. and van Til-
beurgh, H. (2004) J. Biol. Chem., 279(1), 619-625.
[62] Bornemann, S.; Lawson, D. M. and Thorneley, R. N. (2003) Struc-
ture, 11(12), 1463-1465.
[63] Binda, C.; Coda, A.; Angelini, R.; Federico, R.; Ascenzi, P. and
Mattevi. A. (1999) Structure, 7, 265-276.
[64] Yue, Q. K.; Kass, I. J.; Sampson, N.S. and Vrielink, A. (1999)
Biochemistry, 38, 4277-4286.
[65] Barber, M. J.; Neame, P. J.; Lim,L. W.; White, S. and Matthews, F.
S. (1992) J. Biol. Chem., 267, 6611–6619.
[66] Davies, G.M.; Barrett-Bee, K.J.; Jude, D.A.; Lehan, M.; Nichols,
W.W.; Pinder, P.E.; Thain, J.L.; Watkins, W.J. and Wilson, R.G.
(1994) Antimicrob. Agents. Chemother., 38(2), 403-406.
[67] Abeles, R.H. and Alston, T.A. (1990) J. Biol. Chem., 265(28),
16705–16708.
[68] Kollonitsch, J. (1978) Isr. J. Chem. 17, 53-59.
[69] Davis, B. D. and Mingioli, E. S. (1950) J. Bacteriol., 60, 17-28.
[70] Hawkes, T. R.; Lewis, T.; Coggins, J. R.; Mousdale, D. M.; Lowe,
D. J. and Thorneley, R. N. F. (1990) Biochem. J., 265, 899-902.
[71] von Ragué Schleyer.; P. and Kos, A. J. (1983) Tetrahedron, 39,
1141-1150.
[72] Green, J. M. and Nichols, B. P. (1991) J. Biol. Chem., 266, 12971-
12975.
[73] Thomas, M.G.; Lawson, C.; Allanson, N.M.; Leslie, B.W.; Bottom-
ley, J.R.; McBride, A. and Olusanya, O.A. (2003) Bioorg. Med.
Chem. Lett., 13(3), 423-426.
[74] Basso, L. A., Pereira da Silva, L. H., Fett-Neto, A. G., de Azevedo
Jr., W. F., Moreira, I. S., Palma, M. S., Calixto, J. B., Astolfi Filho,
S., dos Santos, R. R., Soares, M. B. P., Santos, D. S. (2005) Mem.
Inst. Oswaldo Cruz, 100(6): 475-506.
[75] da Silveira, N. J. F., Uchoa, H. B., Canduri, F.,Pereira, J. H., Cam-
era Jr., J. C., Basso, L. A., Palma, M. S., Santos, D. S., de Azevedo
Jr., W. F. (2004) Biochem Biophys. Res. Commun.,
322(1), 100–
104.
[76] Ducati, R. D., Basso, L. A., Santos, D. S. (2007) Current Drug
Targets, 8(1) In press.
[77] Dias, M. V. B., Faim, L. M., Vasconcelos, I. B., Oliveira, J. S.,
Basso, L. A., Santos, D. S., De Azevedo, W. F. Acta Crystal-
lographica F. (In press).
[78] De Azevedo, W.F.Jr.; Mueller-Dieckmann, H.J.; Schulze- Gahmen,
U.; Worland, P.J.; Sausville, E.; Kim S.-H. (1996) Proc. Natl.
Acad. Sci. USA., 93(7), 2735-2740.
[79] Canduri, F.; Uchoa, H.B.; de Azevedo, W.F.Jr. (2004) Biochem.
Biophys. Res. Commun., 324(2), 661-666.
[80] De Azevedo, W.F.Jr.; Canduri, F.; Silveira, N.J.F. (2002) Biochem
Biophys. Res. Commun., 293(1), 566-571.
[81] De Azevedo, W.F.Jr.; Leclerc, S.; Meijer, L.; Havlicek, L.; Strnad,
M.; Kim, S.-H. (1997) Eur. J. Biochem., 243, 518-526.
[82] De Azevedo, W. F. Jr., Canduri, F. (2005) Current Computer-
Aided Drug Design, 1, 53-64.
[83] Uchoa, H. B., Jorge, G. E., da Silveira, N. J., Camera J. C., Can-
duri, F., De Azevedo, WF. (2004) Biochem. Biophys. Res. Com-
mun. 325(4),1481–1486.
[84] De Azevedo, W.F.Jr., Gaspar, R.T., Canduri, F.; Camera, J.C.Jr.,
Silveira, N.J.F. (2002) Biochem Biophys. Res. Commun., 297(5),
1154-1158.
[85] da Silveira, N. J. F., Bonalumi, C. A., Uchoa, H. B., Pereira, J. H.,
Canduri, F.,Pereira, J. H., de Azevedo Jr., W. F. (2006) Cell Bio-
chem. Biophys. 44(3), 366-374.
[86] Kim, S.-H.; Schulze-Gahmen, U.; Brandsen, J.; De Azevedo, W. F.
Jr. (1996) Prog. in Cell Cycle Res., 2, 137-145.
Received: December 12, 2005 Accepted: May 26, 2006 Updated: September 11, 2006
135
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