Download PDF
ads:
UNIVERSIDADE FEDERAL DO RIO GRANDE DO SUL
PROGRAMA DE PÓS-GRADUAÇÃO EM CIÊNCIAS BIOLÓGICAS:
BIOQUÍMICA
Chiquimato Desidrogenase de Mycobacterium
tuberculosis: mecanismos cinético e químico da enzima
recombinante
Isabel Osorio da Fonseca
Orientador: Prof. Dr. Luiz Augusto Basso
Tese apresentada ao Programa de Pós-Graduação em Ciências
Biológicas: Bioquímica da Universidade Federal do Rio Grande do Sul
como pré-requisito para obtenção do grau de Doutor.
Porto Alegre, agosto de 2006.
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
Esta tese foi julgada e aprovada para a obtenção do grau de
doutor em Ciências Biológicas: Bioquímica pelo orientador e pela
comissão examinadora no programa de Pós – Graduação em
Ciências Biológicas: Bioquímica.
Orientador:
Prof. Dr. Luiz Augusto Basso
Faculdade de Biociências
Pontifícia Universidade Católica do Rio Grande do Sul
Comissão Examinadora:
Dr. Carlos Alberto Saraiva Gonçalves
Membro do Programa de Pós-Graduação em Ciências Biológicas:
Bioquímica
Universidade Federal do Rio Grande do Sul
Dr. Arthur Germano Fett-Neto
Membro do Programa de Pós-Graduação em Biologia Celular e
Molecular
Universidade Federal do Rio Grande do Sul
Dr. Walter Filgueira de Azevedo Jr.
Departamento de Ciências Fisiológicas da Faculdade de Biociências
Pontifícia Universidade Católica do Rio Grande do Sul
ii
ads:
Este trabalho foi realizado no Centro de Pesquisas em Biologia
Molecular e Funcional, Instituto de Pesquisas Biomédicas,
TECNOPUC, Pontifícia Universidade Católica do Rio Grande do Sul,
sob orientação do Prof. Dr. Luiz Augusto Basso.
iii
Este trabalho é dedicado...
ao meu marido, Kiko, por todo o amor, compreensão, força,
carinho e apoio incondicional.
ao meu filho, Jota, pela compreensão, carinho, paciência e apoio
que me deram forças para vencer mais essa etapa.
A eles, meu MMMUUUUIIIIITTTTOOOO OBRIGADO por me
amarem e estarem ao meu lado...
iv
Agradecimentos
Ao meu orientador Dr. Luiz Augusto Basso, pelos ensinamentos, pela
confiança... sinceramente... MUITO OBRIGADO!!!
Ao Dr. Diógenes Santiago Santos, pela oportunidade, pela confiança, pelos
ensinamentos, pelo exemplo... MUITO OBRIGADO POR TUDO!!!
Ao meu colega Rafael pela enorme força para o desenho dos experimentos,
análise dos resultados, por todos os ensinamentos... MMUUIIITTTTOOOOO
OBRIGADO!!!
A colega Cláudia pelos ensinamentos em bioinformática... MUITO OBRIGADA!!!
Aos meus pais pela força, pelo exemplo, pelo amor, por acreditarem em mim...
AMO MMUUIIITTTTOOOOO VOCÊS!!!
Aos meus colegas, Cristopher, Clarissa, Jotinha, Bruna, Raquel, , Fer,
Isabelinha... pela amizade, pelas contribuições, pelo apoio, pelas brincadeiras,
pelas conversas... que tornaram ainda mais agradáveis meus dias... MUITO
OBRIGADO!!!
À Clotilde, Renilda e Gleci pela dedicação zelando sempre a manutenção e
funcionamento do laboratório... MUITO OBRIGADO!!!
Ao Programa de Pós-Graduação em Ciências Biológicas: Bioquímica da
Universidade Federal do Rio Grande do Sul pela oportunidade que me foi dada.
À CAPES pelo apoio financeiro que me auxiliou nesse período.
v
Índice
Parte I 1
Resumo 2
Abstract 3
Lista de Abreviaturas 4
Introdução 5
Tuberculose 5
Via do Chiquimato e Chiquimato desidrogenase 11
Objetivo 19
Parte II 21
Capítulo I: 22
Functional shikimate dehydrogenase from Mycobacterium tuberculosis
H37Rv: purification and characterization.
Capítulo II: 23
Shikimate Dehydrogenase from Mycobacterium tuberculosis H37Rv:
Kinetic and Chemical Mechanisms.
Parte III 24
Discussão
Capítulo I: 25
Functional shikimate dehydrogenase from Mycobacterium tuberculosis
H37Rv: purification and characterization.
Capítulo II: 35
Shikimate Dehydrogenase from Mycobacterium tuberculosis H37Rv:
Kinetic and Chemical Mechanisms.
Conclusão 50
Referências 51
vi
Anexos
Anexo I 55
1. AZEVEDO, WF; SILVA, RG; FONSECA, IO; RENARD, G; BASSO, LA;
SANTOS, DS. (2003) Docking and small angle X-ray scattering studies of
purine nucleoside phosphorylase. Biochemical and Biophysical Research
Communications 309, 928-933.
Anexo II 56
2. POLETTO, SS; FONSECA, IO; CARVALHO, LP; BASSO, LA; SANTOS, DS.
(2004) Selection of an Escherichia coli host that expresses mutant forms
of Mycobacterium tuberculosis 2-trans enoyl-ACP(CoA) reductase and 3-
Ketoacyl-ACP (CoA) reductase enzymes. Protein Expression and Purification
34, 118-125.
Anexo III 57
3. RIZZI, C; FONSECA, I.O.; GALLAS, M; WEBER, P; FRAZZON, J; SANTOS,
DS; BASSO, LA. (2005) DAHP synthase of Mycobacterium tuberculosis
H37Rv: Cloning, overexpression, and purification of the functional
enzyme. Protein Expression and Purification 40, 23-30.
Anexo IV 58
4. PATENTE (2005) SANTOS, D. S.; BASSO, L.A.; RENARD, G.; FONSECA,
I.O.; CHIES, J.M. Método de Obtenção de Seqüências Nucleotídicas
Quiméricas e Seqüência Nucleotídica Quimérica - INPI Registro PI0506047-
8.
Anexo V 59
5. LEOPOLDINO, A.M.; CANDURI, F.; CABRAL, H.; JUNQUEIRA, M.;
MARQUI, A.B.T.; APPONI, L.H.; FONSECA, I.O.; DOMONT, G.B.; SANTOS,
D. S.; VALENTINI, S. (2006) Expression, purification, and circular
dichroism analysis of human CDK9. Protein Expression and Purification 47,
614-620.
vii
PARTE I
1
Resumo
O aumento na prevalência da tuberculose (TB), a emergência de cepas resistentes a múltiplas
drogas de Mycobacterium tuberculosis, o agente etiológico da TB, e o efeito devastador da co-
infecção pelo HIV têm enfatizado a urgente necessidade no desenvolvimento de novos agentes
antimicobacterianos. A análise completa da seqüência genômica do M. tuberculosis H37Rv
mostrou a existência de genes envolvidos na via de biossíntese de aminoácidos aromáticos, a
via do chiquimato, e evidências experimentais indicam que esta via é essencial para o M.
tuberculosis e apresenta-se ausente no homem. Os produtos dos genes que são essenciais
para o crescimento dos microrganismos fazem deles alvos atrativos para a ação de drogas,
pois a inibição de sua função pode matar o bacilo. Previamente, nosso grupo relatou a
clonagem do gene aroE de M. tuberculosis e a expressão do seu produto na forma solúvel, a
enzima chiquimato desidrogenase (MtbSD), que catalisa o quarto passo na via do chiquimato.
Neste trabalho apresentamos, no primeiro artigo, intitulado “Functional shikimate
dehydrogenase from Mycobacterium tuberculosis H37Rv: purification and characterization”, a
purificação da MtbSD solúvel e ativa por cromatografia líquida, o seqüenciamento do N-
terminal, a espectrometria de massas, a determinação do estado oligomérico por cromatografia
em gel filtração, a estabilidade térmica, a determinação de parâmetros cinéticos aparentes em
estado estacionário para as reações direta e reversa, a constante de equilíbrio aparente e
energia de ativação para a reação química catalisada pela MtbSD. No segundo artigo, intitulado
Shikimate Dehydrogenase from Mycobacterium tuberculosis H37Rv: Kinetic and Chemical
Mechanisms”, descrevemos os parâmetros de velocidade inicial na reação direta, os estudos de
inibição pelos produtos, os efeitos isotópicos cinéticos primários do deutério, os efeitos
isotópicos cinéticos do solvente, os efeitos isotópicos cinéticos múltiplos, proton inventory, o
efeito de pH na química ácido/base para a ligação e catálise dos substratos, e a estrutura 3D da
MtbSD obtida in silicon pela modelagem por homologia. Esses resultados servem como base
para os estudos cinéticos e estruturais que podem auxiliar no desenho racional de inibidores a
serem testados como agentes antimicobacterianos.
2
Abstract
The increasing prevalence of tuberculosis (TB), the emergence of multidrug-resistant strains of
Mycobacterium tuberculosis, the causative agent of TB, and the devastating effect of co-
infection with HIV have highlighted the urgent need for the development of new
antimycobacterial agents. Analysis of the complete genome sequence of M. tuberculosis H37Rv
shows the presence of genes involved in the aromatic amino acid biosynthetic pathway, the
shikimate pathway, and experimental evidence showed that this pathway is essential for M.
tuberculosis and it is absent in humans. The gene products that are essential for the growth of
the microorganisms make them attractive drug targets since inhibiting their function may kill the
bacilli. Our group has previously reported the cloning of M. tuberculosis aroE gene and the
expression of the product in the soluble form, the shikimate dehydrogenase (MtbSD) enzyme,
that catalysis the fourth reaction in the shikimate pathway. Currently, in the first manuscript
Functional shikimate dehydrogenase from Mycobacterium tuberculosis H37Rv: purification and
characterization”, we present the purification of soluble and active MtbSD, N-terminal
sequencing, mass spectrometry, assessment of the oligomeric state by gel filtration
chromatography, thermal stability, determination of apparent steady-state kinetic parameters for
both forward and reverse directions, apparent equilibrium constant, and energy of activation for
the enzyme-catalyzed chemical reaction. In the second manuscript “Shikimate Dehydrogenase
from Mycobacterium tuberculosis H37Rv: Kinetic and Chemical Mechanisms”, we describe the
kinetic mechanism, initial velocity patterns in the forward direction, product inhibition studies,
primary deuterium kinetic isotope effects, solvent kinetic isotope effects, multiple isotope effects,
proton inventory, pH rate profile and the MtbSD 3D structure obtained in silicon by homology
modeling. These results should be useful as a solid base for structural and kinetic studies, which
can aid in the rational design of inhibitors to be tested as antimycobacterial agents.
3
Lista de Abreviaturas
DAHP
3-desoxi-D-arabinoheptulosonato-7-fosfato
DHS
3-desidrochiquimato
EcoliQSD quinato-chiquimato desidrogenase de Escherichia coli
EcoliSD chiquimato desidrogenase de Escherichia coli
EPSP 5-enolpiruvilchiquimato-3-fosfato
E4P eritrose-4-fosfato
HinflSD chiquimato desidrogenase de Haemophilus influenzae
IPTG β-D-tiogalactopiranosídeo
MjanSD chiquimato desidrogenase de Methanacoccus jannaschii
MtbSD chiquimato desidrogenase de Mycobacterium tuberculosis
PEP Fosfoenolpiruvato
PsatSD chiquimato desidrogenase de Pisum sativum
SD chiquimato desidrogenase
SHK
chiquimato ou ácido chiquímico
TB tuberculose
4
Introdução
Tuberculose
A tuberculose (TB) é a doença infecto-contagiosa responsável pela maior parte
da mortalidade humana causada por um único agente infeccioso, o Mycobacterium
tuberculosis (ENARSON & MURRAY, 1996).
No século XIX, a TB já era uma doença devastadora, com taxa de transmissão e
índice de morbidade altíssimos; período que corresponde ao começo da revolução
industrial, que trouxe como conseqüência a formação de aglomerados urbanos com
péssimas condições de higiene e habitação (DORMANDY, 2000).
A infecção ocorre por meio da inalação do bacilo, presente em partículas
infectivas, expelidas por pacientes com tuberculose ativa através da tosse ou espirro.
Elas apresentam dimensão suficiente para permanecerem suspensas no ar de
ambientes fechados e assim alcançar os alvéolos pulmonares de outros indivíduos
(DUNLAP e cols., 2000; GLICKMAN & JACOBS, 2001).
Após a infecção, a doença é regulada fundamentalmente pelo sistema imune do
hospedeiro, que determina se o microrganismo é eliminado, condicionado a um estado
de latência, ou se está propenso a desenvolver a doença ativa no seu hospedeiro.
Assim que a micobactéria chega aos bronquíolos e alvéolos, os macrófagos
fagocitam o invasor. A partir daí, a infecção irá ou não se estabelecer dependendo da
virulência bacteriana e da capacidade bactericida dos macrófagos do hospedeiro. Caso
a micobactéria não seja eliminada pela primeira linha de defesa do organismo, a
erosão da bactéria no interstício pulmonar estabelecerá uma inflamação local; logo, o
dano tissular, a formação dos granulomas e a agregação destes formarão o tubérculo -
5
a característica nodular da lesão de TB (MILBURN, 2001). Aproximadamente 4 meses
após a infecção, ocorre a eliminação da maior parte dos bacilos e cessa a infecção
primária.
A tuberculose conhecida como tuberculose extra-pulmonar ou miliar, representa
15% dos pacientes com a infecção ativa. Ela é causada pelo crescimento bacteriano
excessivo no granuloma, que atinge então a corrente sangüínea e disseminando-se
para a pleura, linfonodos, fígado, baço, ossos e articulações, coração, cérebro, sistema
genito-urinário, meninges, peritônio ou na pele.
Uma outra possibilidade é a coexistência pacífica entre invasor e hospedeiro
humano, que recebe a denominação de infecção latente. Nesse caso se estabelece um
reservatório bacteriano no indivíduo infectado, onde o metabolismo bacteriano está
reduzido como conseqüência da ação do sistema imunológico do hospedeiro. No
momento em que o sistema imune do hospedeiro apresentar falhas, como subnutrição,
câncer, diabetes, quimioterapia, entre outros, pode ocorrer a reativação do bacilo,
levando o indivíduo a desenvolver TB ativa (PARRISH e col., 1998).
O M. tuberculosis adapta-se a uma variedade de condições ambientais, dentro e
fora do hospedeiro, sendo então considerado o patógeno de maior sucesso da história.
Esta característica peculiar do bacilo faz com que a TB permaneça como uma das
principais causas de morte no mundo (ENSERINK, 2001). A exata forma como o
microrganismo permanece silencioso no corpo humano, em trégua com o sistema
imune, tem sido um longo mistério (WICKELGREN, 2000).
A TB começou a ser esclarecida no meio científico em 1882, quando no IV
Congresso Mundial de TB, o médico e bacteriologista alemão Robert Koch tornou
6
pública a identificação do agente etiológico da doença. Este marco da história científica
tornou possível a procura de terapias eficazes no tratamento de pacientes com TB.
A descoberta da estreptomicina, em 1944, marcou a era de ouro no
desenvolvimento de drogas para TB. Outras drogas, ainda hoje utilizadas para o
tratamento da doença, foram sendo introduzidas: o ácido p-aminosalicílico (1946),
isoniazida (1952), ciclosserina (1955), canamicina (1957), rifampicina (1965),
etionamida (1966), etambutol (1968) e pirazinamida (1970) (DUNCAN, 2003). A
administração de estreptomicina, isoniazida e ácido para-amino-salicílico contra a
doença ativa reduziu consideravelmente a mortalidade por TB (BLOOM & MURRAY,
1992; DANIEL, 1997). Essas drogas e as vacinas de prevenção da infecção trouxeram
um período de relativa tranqüilidade com respeito à doença.
Em 1993, a TB foi declarada “emergência de saúde global” pela Organização
Mundial da Saúde (OMS). Estima-se que anualmente ocorram 8 milhões de novos
casos e 3 milhões de mortes devido a esta doença (DUNLAP e cols., 2000). Em torno
de um terço da população mundial está hoje infectada pelo M. tuberculosis e 10%
deste número desenvolverá a doença ao longo da vida (ENSERINK, 2001). Além disso,
a OMS e o Centro de Controle e Prevenção de Doenças dos Estados Unidos indicam a
TB como responsável por 27% de todas as mortes evitáveis em todo o mundo
(SEPKOWITZ e cols., 1995). Estima-se que até 2020, um bilhão de pessoas se
tornarão infectadas, 200 milhões de pessoas desenvolverão a doença e 70 milhões
morrerão devido à TB se a atual tendência permanecer inalterada (PASQUALOTO &
FERREIRA, 2001).
7
A argumentação de que a tuberculose é uma doença reemergente é válida
apenas para alguns países europeus e para os Estados Unidos; pois no Brasil, a TB
não é um problema reemergente e sim um problema persistente (RUFFINO-NETTO,
2002).
Os fatores responsáveis pelo ressurgimento da TB, até então considerada
vencida em países desenvolvidos são: (1) advento da Síndrome da Imunodeficiência
Adquirida (AIDS); (2) deterioração dos programas de saúde pública visando o controle
da TB; (3) aumento da transmissão do M. tuberculosis em hospitais e prisões; e (4)
surgimento de cepas resistentes a múltiplas drogas (MDR, resistentes a pelo menos
isoniazida e rifampicina) (ESPINAL, 2003).
A TB e a AIDS são referenciadas como co-epidemia, visto que nesta situação a
probabilidade de desenvolvimento da doença aumenta até 30 vezes. A taxa de
reativação de TB latente em pacientes HIV-positivos é de 2,3 a 13,3%, dependendo do
nível de imunossupressão, contra 0,2% em pessoas imunocompetentes (KNIGGE e
cols., 2000). Pacientes HIV-positivos infectados com M. tuberculosis progridem para TB
ativa em uma taxa de 37% nos primeiros 6 meses, ao contrário de pacientes
imunocompetentes cuja taxa é de 2 a 5% nos primeiros dois anos (SEPKOWITZ,
1995). Esse aumento considerável na probabilidade leva a uma aceleração da
transmissão do agente etiológico (KNIGGE e cols., 2000) e, portanto, a co-infecção TB
- HIV representa um problema de efeito devastador tanto para pacientes infectados
como para a população em geral.
O aumento da transmissão do M. tuberculosis em hospitais e prisões ocorre
devido aos aglomerados humanos com condições de higiene e habitação péssimas.
8
Um agravante deste quadro decorre da deterioração dos sistemas de saúde visando o
controle da TB (FÄTKENHEUER e col., 1999).
A TB resistente a múltiplas drogas (MDR-TB – “multidrug-resistant tuberculosis”)
está relacionada à resistência aos medicamentos para combater a doença, visto que
ocorre a resistência à pelo menos rifampicina e isoniazida. Em 2000 foi estimado que
3,2% dos novos casos são de MDR-TB (Espinal, 2003). Segundo a OMS, ocorrem
300.000 novos casos de MDR-TB anualmente, onde 79% são de “super cepas”, que se
refere às cepas resistentes a pelo menos três das quatro principais drogas usadas no
tratamento de TB (WHO, 2004).
Através de um estudo realizado em 35 países pela OMS e a União Internacional
Contra a Tuberculose e Doenças Pulmonares (IUATLD – “International Union Against
Tuberculosis and Lung Disease”) foi demonstrado que a doença consiste em um
problema global, pois todas as regiões analisadas apresentaram linhagens de M.
tuberculosis com resistência à pelo menos uma das drogas antituberculose, sendo
normalmente a isoniazida ou estreptomicina (PABLOS-MENDEZ e col., 1998).
O MDR-TB surge sob pressão seletiva da quimioterapia inapropriada ou da não
adesão ao tratamento (RILEY, 1993). Terapia inadequada compreende o uso de uma
única droga, combinações inadequadas de drogas, má-absorção de drogas
administradas, etc. Nesses casos o M. tuberculosis estará exposto a concentrações
sub-letais de compostos bactericidas, quando as cepas mais resistentes na população
bacteriana sobreviverão e superarão a cepa sensível (PETRINI & HOFFNER, 1999).
No caso da co-epidemia de HIV-TB, o aparecimento das linhagens MDR-TB foi
favorecido pelo maior índice de abandono do tratamento (BRENNAN, 1997).
9
Atualmente o tratamento recomendado pela OMS consiste na administração
combinada dos antibióticos isoniazida, rifampicina, pirazinamida e estreptomicina (ou
etambutol) durante 2 meses, seguida pela combinação de isoniazida e rifampicina por
pelo menos mais 4 meses.
Mesmo com um tratamento bem estabelecido, a eficácia do tratamento contra o
bacilo de Koch através da utilização de antibióticos tem sido dificultada, em parte,
principalmente pela oposição do paciente em cumprir com o tratamento de 6 meses
(YOUNG, 1998). A longa duração e os desagradáveis efeitos colaterais do tratamento
levaram a OMS ao desenvolvimento de um programa conhecido como DOTS (directly
observed treatment short-course) que visa acompanhar a adesão dos pacientes ao
tratamento, monitorar seu progresso e observar a ingestão de cada dose da medicação
(NSB Editorial comment, 2000).
O programa DOTS combina cinco elementos: compromisso político, serviços de
microscopia, suprimento de medicação, sistemas de monitoramento e observação
direta do tratamento (PASQUALOTO & FERREIRA, 2001). Entretanto, em alguns
países super populosos o programa não consegue alcançar 100% dos indivíduos
infectados, não chegando aos objetivos primeiramente propostos. Uma alternativa é a
obtenção de drogas que simplifiquem o tratamento da TB facilitando a execução do
programa DOTS (DUNCAN, 2003).
Desenvolvimento de Agentes anti-TB
Frente a essa exposição de fatos que envolvem a TB, faz-se necessária à
introdução de novas drogas anti-TB, pois desde a metade da década de 70 não houve
a descoberta, nem tampouco o desenvolvimento de novas drogas de primeira linha que
10
possam ser utilizadas no combate ao bacilo causador da TB (PETRINI & HOFFNER,
1999).
As novas drogas teriam como objetivo a efetividade no tratamento de MDR-TB e
TB latente, além da redução no tempo do tratamento (O´BRIEN & NUNN, 2001). Isto
representaria vantagem significativa em relação às drogas atualmente utilizadas.
DUNCAN (2003) estima que o desenvolvimento completo de uma nova droga anti-TB
custe de 100 a 800 milhões de dólares. Considerando-se que 95% dos novos casos de
TB ocorram em países em desenvolvimento, poucos pacientes poderiam pagar os altos
custos necessários para o lucro da indústria farmacêutica (O´BRIEN & NUNN, 2001). O
provável retorno financeiro é baixo e, conseqüentemente, o interesse neste setor
também é baixo (DUNCAN, 2003).
Mais de um século após a descoberta de Robert Koch, está disponível uma nova
perspectiva para o desenvolvimento de novas drogas, em decorrência da seqüência
completa do genoma do M. tuberculosis, marcando uma nova era na batalha contra
esse patógeno (COLE e cols., 1998).
Utilizando a análise de homologia de seqüências, foi possível a identificação de
genes estruturais envolvidos em rotas metabólicas. A identificação e validação de rotas
essenciais ao microrganismo, ausentes no hospedeiro, são o primeiro passo no
desenvolvimento de inibidores específicos de baixa toxicidade (DUNCAN, 2003).
Alguns genes expressos na infecção latente foram descritos, e suas proteínas podem
ser bons alvos para o desenvolvimento de drogas para o tratamento da TB latente
(WICKELGREN, 2000).
11
Uma possibilidade para a obtenção de novas drogas é o desenho racional que
pode ser alcançado quando existe informação detalhada sobre a interação entre
proteína alvo - ligante, fornecida pela estrutura tridimensional do alvo (DUNCAN, 2003).
Com a obtenção da estrutura 3D do alvo é possível fazer uma triagem em biblioteca
virtual de compostos químicos que possam ser utilizados como drogas naturais ou que
sirvam como base para o desenho racional de novas drogas sintéticas. As informações
sobre os mecanismos cinético e químico também são importantes para o desenho
racional de inibidores.
Uma alternativa atraente para obter novos inibidores específicos para alvos
definidos é a busca desses compostos na biodiversidade brasileira, e para isso é
necessária à obtenção de um alvo isolado e funcional. Utilizando a tecnologia de
detecção através da Ressonância Plasmônica de Superfície (SPR), imobiliza-se o alvo
e após faz-se a triagem de ligantes (possíveis inibidores) presentes em extratos de
plantas oriundos da biodiversidade brasileira e/ou seus compostos purificados. Essa
tecnologia monitora as interações bimoleculares entre compostos e o alvo de interesse
em tempo real, detectando a ligação de compostos ao alvo pelo aumento de massa em
relação ao estado inicial. Esse tipo de análise e metodologia é ideal para a descoberta
de novas drogas por ter alta sensibilidade, e pelo fato de que os compostos químicos
analisados não necessitam de marcação radioativa nem de fluorescência. Em posse
dos possíveis inibidores, segue-se para a determinação das constantes de inibição (K
i
)
para a caracterização da natureza da inibição (BASSO e col., 2005).
Dentre as vias metabólicas identificadas no genoma do M. tuberculosis está a
via do chiquimato (Figura 1) que representa um alvo validado para o desenvolvimento
12
de herbicidas e agentes antimicrobianos, pois estão presentes em algas, vegetais,
bactérias, fungos (BENTLEY, 1990) e parasitas do filo Apicomplexa (ROBERTS e cols.,
1998), mas ausente em vertebrados. Foi sugerido que esta via é essencial para o
crescimento do M. tuberculosis mesmo na presença de suplementos exógenos
(PARISH & STOKER, 2002). Com isso, inibidores específicos das enzimas presentes
nessa via apresentam potencial como agentes terapêuticos no tratamento da TB.
A via do chiquimato é composta de uma seqüência de sete reações que vão
desde a condensação da D-eritrose 4-fosfato (E4P) e do fosfoenolpiruvato (PEP) até a
formação do corismato. O corismato leva à biossíntese de precursores de aminoácidos
aromáticos, vitaminas E e K, ácido fólico (BENTLEY, 1990), naftoquinonas,
menaquinonas e micobactinas (em micobatérias) (Figura 2) (RATLEDGE, 1982).
13
Figura 1: A Via do Chiquimato. Esta via é composta por sete etapas enzimáticas que
culminam com a formação do corismato, o precursor chave para formação de
compostos aromáticos essenciais aos microrganismos.
14
Figura 2: Importância da Via do Chiquimato. O produto da via do chiquimato,
corismato e os compostos essenciais sintetizados a partir dele.
15
O trabalho aqui descrito mostra o estudo sobre os mecanismos cinético e químico
feito com um alvo enzimático, a chiquimato desidrogenase de M. tuberculosis (MtbSD),
presente na quarta etapa de catálise da via do chiquimato (Figura 1). A MtbSD catalisa
a redução reversível do 3-desidrochiquimato (DHS) a chiquimato (SHK), utilizando
especificamente o cofator NADPH
(Figura 3) (BENTLEY, 1990; CHAUDHURI e cols.,
1985).
chiquimato
desidrogenase
Figura 3: Reação enzimática catalisada pela enzima chiquimato desidrogenase. A
redução do 3-desidrochiquimato dependente de NADPH leva à formação do D-
chiquimato.
A proteína MtbSD, codificada pelo gene aroE (Rv2552c), possui 269
aminoácidos e pertence à superfamília das oxiredutases dependentes de NAD(P)H,
que atuam em rotas anabólicas e catabólicas (BENACH e cols., 2003).
Em 2002, a primeira descrição sobre a enzima MtbSD foi publicada
(MAGALHÃES e cols.). O artigo trata da clonagem do gene aroE da cepa M.
tuberculosis H37Rv em vetor para expressão de proteínas heterólogas pET23a(+)
(Novagen
®
) e expressão da enzima recombinante em E. coli BL21(DE3). O trabalho
16
relatou que o sistema pET23a(+)-aroE em E. coli BL21(DE3) é excelente para a
expressão do alvo MtbSD recombinante sem adição do indutor IPTG e a obtenção da
MtbSD recombinante solúvel e funcional foi feita por meio da utilização do
procedimento de congelamento e descongelamento em ciclos repetidos da E. coli
BL21(DE3) com pET23a(+)-aroE para o extravasamento do conteúdo celular
(MAGALHÃES e cols.).
A expressão da MtbSD também foi obtida em vetor de expressão pET28b
(Novagen
®
). A purificação da MtbSD foi feita em coluna de níquel por meio de ligação
da cauda de histidina da proteína recombinante. Os valores de K
m
aparente para os
substratos ácido chiquímico e NADP
+
foram de 0,030 mM e 0,063 mM,
respectivamente, e o k
cat
foi de 399 s
-1
. Esses valores foram determinados sem retirar
da cauda de histina da MtbSD recombinante, o que pode influenciar os dados obtidos
(ZHANG e cols., 2005).
A estrutura da chiquimato desidrogenase codificada pelo gene aroE de E. coli
(EcoliSD) foi determinada com resolução de 1,5 Å. A enzima apresenta uma arquitetura
com dois domínios
α
/
β
separados por um sulco grande que corresponde ao sítio
catalítico da enzima (MICHEL e col., 2003). O sítio ativo da EcoliSD foi identificado
pelos resíduos Ser14, Ser16, Lys65, Asn86, Thr101, Asp102 e Gln244, conservados na
família das chiquimato desidrogenase. A caracterização cinética aparente da proteína
EcoliSD, determinação de K
m
e k
cat
, foi feita utilizando a reação reversa que
corresponde à catálise de oxidação do ácido chiquímico. A EcoliSD trata-se de um
monômero em solução (CHAUDURI & COGGINS, 1985).
17
Um estudo semelhante foi feito com a chiquimato desidrogenase codificada pelo
gene aroE de Methanococcus jannaschii (MjanSD). A estrutura foi determinada em
2,35 Å e a arquitetura descrita é semelhante à EcoliSD. A proteína MjanSD apresenta-
se na forma dimérica em solução (PADYANA e BURLEY, 2003).
18
Objetivo
Considerando que a tuberculose é problema de saúde pública global, o quadro
alarmante com o advento da AIDS, a inexistência de programas eficazes de saúde
publica com relação à TB, o surgimento de cepas resistentes, o interesse na obtenção
de novos antimicobacterianos, a via do chiquimato é um alvo promissor, e que
inibidores específicos das enzimas presentes nessa via apresentam potencial como
agentes terapêuticos no tratamento da TB, “o presente trabalho visa elucidar os
mecanismos cinético e químico da enzima chiquimato desidrogenase de M.
tuberculosis (MtbSD), que servirão como base para o alcance de um inibidor
específico com possível utilização no tratamento da tuberculose.”
Para atingir tal objetivo foram feitas:
9 Caracterização da MtbSD
Determinação dos parâmetros cinéticos aparentes em ambas as direções da
reação, da constante de equilíbrio da reação, da estabilidade térmica da enzima
recombinante e da energia de ativação para a reação química.
9 Determinação dos mecanismos cinético e químico para a MtbSD
Com o intuito de investigar os mecanismos cinético e químico foram feitos
experimentos de velocidade inicial em estado estacionário, de estudos de
inibição pelos produtos, de efeitos isotópicos cinéticos primário, de efeitos
isotópicos cinéticos de solvente, de efeitos isotópicos cinéticos múltiplos, de
proton inventory” e de efeito de pH na química ácido/base na reação.
19
9 Estudos estruturais da MtbSD utilizando modelagem por homologia
Os estudos estruturais in silicon da MtbSD são úteis na interpretação dos
resultados obtidos nos experimentos cinéticos e possibilita a identificação dos
possíveis resíduos envolvidos na ligação e catálise dos substratos. Para tal foi
modelada a estrutura 3D para a MtbSD em silicon utilizando como molde a
chiquimato desidrogenase de E. coli.
20
PARTE II
21
Capítulo I
Fonseca, I. O., Magalhães, M. L. B., Oliveira, J. S., Silva, R. G.,
Mendes, M. A., Palma, M. S., Santos, D. S., Basso, L. A. (2006)
Functional shikimate dehydrogenase from Mycobacterium
tuberculosis H37Rv: purification and characterizationProtein Expr.
Purif. 46, 429-437.
Arquivo nomeado:
Functional shikimate dehydrogenase from Mycobacterium tuberculosis H37Rv.pdf
22
0
Functional shikimate dehydrogenase from Mycobacterium
tuberculosis H37Rv: Purification and characterization
Isabel O. Fonseca
a
, Maria L.B. Magalha
˜
es
a
, Jaim S. Oliveira
a
, Rafael G. Silva
a
,
Maria A. Mendes
b
, Mario S. Palma
b
, Dio
´
genes S. Santos
a,
*
, Luiz A. Basso
a,
*
a
Centro de Pequisas em Biologia Molecular e Funcional, TECNOPUC, Programa de Po
´
s-Graduac¸a
˜
o em Cie
ˆ
ncias Biolo
´
gicas: Bioquı
´mica,
Programa de Po
´
s-Graduac¸a
˜
o em Biologia Celular e Molecular (PUCRS-UFRGS), Pontifı
´c
ia Universidade Cato
´
lica do Rio Grande do Sul,
Porto Alegre—RS 90619-900, Brazil
b
Laborato
´
rio de Biologia Estrutural e Zooquı
´
mica, Centro de Estudos de Insetos Sociais, Departamento de Biologia, Instituto de Biocie
ˆ
ncias,
Universidade Estadual Paulista, Rio Claro—SP 13506-900, Brazil
Received 18 July 2005, and in revised form 30 September 2005
Available online 27 October 2005
Abstract
Tuberculosis (TB) poses a major worldwide public health problem. The increasing prevalence of TB, the emergence of multi-drug-
resistant strains of Mycobacterium tuberculosis, the causative agent of TB, and the devastating effect of co-infection with HIV have
highlighted the urgent need for the development of new antimycobacterial agents. Analysis of the complete genome sequence of M. tuber-
culosis shows the presence of genes involved in the aromatic amino acid biosynthetic pathway. Experimental evidence that this pathway is
essential for M. tuberculosis has been reported. The genes and pathways that are essential for the growth of the microorganisms make
them attractive drug targets since inhibiting their function may kill the bacilli. We have previously cloned and expressed in the soluble
form the fourth shikimate pathway enzyme of the M. tuberculosis, the aroE-encoded shikimate dehydrogenase (mtSD). Here, we present
the purification of active recombinant aroE -encoded M. tuberculosis shikimate dehydrogenase (mtSD) to homogeneity, N-terminal
sequencing, mass spectrometry, assessment of the oligomeric state by gel filtration chromatography, determination of apparent
steady-state kinetic parameters for both the forward and reverse directions, apparent equilibrium constant, thermal stability, and energy
of activation for the enzyme-catalyzed chemical reaction. These results pave the way for structural and kinetic studies, which should aid
in the rational design of mtSD inhibitors to be tested as antimycobacterial agents.
Ó 2005 Elsevier Inc. All rights reserved.
Keywords: Mycobacterium tuberculosis; Shikimate pathway; Shikimate dehydrogenase; aroE; Enzyme kinetics; Protein purification; Drug target
The causative agent of tuberculosis (TB),
1
Mycobacteri-
um tuberculosis, infects approximately 32% of the worldÕs
human population. TB remains the leading cause of mor-
tality due to a bacter ial pathogen. Currently, there are 8
million new cases and 2 million deaths annually from
tuberculosis, and it is predicted that a total of 225 million
new cases and 79 million deaths will occur between 1998
and 2030 [1]. Approximately 2 billion individuals are
believed to harbor latent TB based on tuberculin skin test
surveys [2], which represents a considerable reservoir of
bacilli. Possible factors underlying the resurgence of TB
worldwide include the HIV epidemic, increase in the home-
less population, and decline in health care structures and
national surveillance [3]. The pandemic of AIDS has had
1046-5928/$ - see front matter Ó 2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.pep.2005.10.004
*
Corresponding authors. Fax: +55 51 33203629.
E-mail addresses: [email protected] (D.S. Santos), luiz.basso@
pucrs.br (L.A. Basso).
1
Abbreviations used: DHS, 3-dehydroshikimate; ESI-MS, electrospray
ionization mass spectrometry; IPTG, isopropyl b-
D-thiogalactopyrano-
side; LB, Luria–Bertani; MDR-TB, multi-drug-resistant tuberculosis;
mtSD, Mycobacterium tuberculosis shikimate dehydrogenase; NADP
+
,
oxidized b-nicotinamide adenine dinucleotide phosphate; NADPH,
reduced b-nicotinamide adenine dinucleotide phosphate; SHK,
D-shikim-
ate; SDS–PAGE, sodium dodecylsulfate–polyacrylamide gel electropho-
resis; TB, tuberculosis.
www.elsevier.com/locate/yprep
Protein Expression and Purification 46 (2006) 429–437
a major impact on the TB problem, owing not only to
increased reactivation of latent TB but also to acceleration
of transmission in HIV/TB co-infected patients following
the increase in the number of smear-positive infectious pul-
monary TB cases [4]. Another contributing factor is the
evolution of multi- drug TB (MDR-TB), defined as resis-
tant to isoniazid and rifampicin, which are the most effec-
tive first-line drugs [5]. MDR-TB is more difficult and
more expensive to treat, and more likely to be fatal [6].
According to the 2004 Global TB Control Report of the
World Health Orga nization, there are 300,000 new cases
per year of MDR -TB worldwide, and 79% of MDR-TB
cases are now ‘‘super-strains,’’ resistant to at least three
of the four main drugs used to treat TB [7]. The factors that
most influence the emergence of MDR-TB strains include
inappropriate treatment regimens, and patient non-compli-
ance in completing the prescribed courses of therapy due to
the lengthy standard ‘‘short-course’’ treatment or when the
side effects become unbearable [8]. Thus, there is a need for
the development of new antimycobacterial agents to both
treat M. tuberculosis strains resistant to existing drugs
and shorten the duration of short-course treatment to
improve patient compliance [9].
The shikimate pathway is an attractive target for the
development of herbicides and antimicrobial agents
because it is essential in algae, higher plants, bacteria,
and fungi, but absent from mammals [10]. The mycobacte-
rial shikimate pathway leads to the biosynthesis of choris-
mic acid, which is a precursor of aromatic amino acids,
naphthoquinones, menaquinones, and mycobactins [11].
The salicylate-derived mycobactin siderophores have been
shown to be essential for M. tuberculosis growth in macro-
phages [12]. In addition, the shikimate pathway has more
recently been shown to be essential for the viability of
M. tuberculosis [13]. Accordingly, the essentiality of myco -
bacterial shikimate pathway and its absence from human
host indicate that any enzyme of this pathway represents
a promising target for the developm ent of non-toxi c anti-
mycobacterial agents.
Analysis of the complete genome sequence of M. tuber-
culosis shows the presence of seven aro genes predicted to
be involved in the shikimate pathway [14]. Amongst them,
the aroE-encoded shikimate dehydrogenase (SD) has been
predicted by DNA sequence homology to be present in
M. tuberculosis H37Rv strain. Shikimate dehydrogenase
(EC 1.1.1.25) catalyzes the fourth reaction in the shikimate
pathway. We have previously reported the cloning and
expression of M. tuberculosis SD (mtSD) [15]. In addition,
measurements of the NADPH-dependent redu ction of
3-dehydroshikimate to shikimate catalyzed by mtSD con-
firmed the correct assignment to the structural ge ne encod-
ing SD in M. tuberculosis [15]. Here, we report the
purification to homogeneity of recombinant and functional
mtSD. The purification protocol yielded approximately
11 mg of homogeneous recombinant mtSD from 14 L of
Escherichia coli cell culture. We also present N-terminal
amino acid sequencing and electrospray ionization mass
spectrometry (ESI-MS) data that unambiguously demon-
strate the identity and purity of homogenous recombinant
mtSD protein. The estimated molecular mass of native
homogeneous recombinant protein determined by gel fil-
tration indicates that mtSD enzyme is a dimer in solution
with a subunit molecular mass value of 27,207 Da deter-
mined by ESI-MS. The apparent kinetic parameters for
mtSD were determined for all substrates in both forward
and reverse reactions. The mtSD thermal stability was eval-
uated, and an estimate for the activation energy (E
a
) was
obtained from a linear plot of log k versus 1/T (K
À1
). A
comparison of polypeptide sequences of SDs from
M. tuberculosis, E. coli, Haemophilus influenzae,and
Methanococcus jannaschii allowed identification of amino
acid residues that are likely to be involved in 3-dehydros-
hikimate/shikimate binding.
The results presented here will pave the way for structur-
al and functional efforts currently underway in our labora-
tory. It is hoped that these studies will provide a framew ork
on which to base the design of new agents with antituber-
cular activity and low toxicity to humans.
Materials and methods
Overexpression and release of mtSD
The recombinant plasmid pET23a(+)::aroE was trans-
formed into E. coli BL21 (DE3) host cells by electropor-
ation, and selected on LB agar plates containing
50 lgmL
À1
carbenicillin. Single colonies were used to
inoculate 14 L of LB medium containing 50 lgmL
À1
car-
benicillin with no addition of isopropyl b-
D-thiogalacto-
pyranoside (IPTG), an d grown for 24 h at 37 °Cat
180 rpm as described elsewhere [15]. Cells (49 g) were
harvested by centrifugation at 14,900g, for 30 min at
4 °C, and stored at À20 °C. Cells expressing recombinant
mtSD were placed into metal containers to allow fast
temperature equilibrium to be reached, which is neces-
sary for increased efficiency of cell disruption by the
freeze–thaw method [15]. Cell paste was placed into a
dry-ice/ethanol bath for 2 min and immediately trans-
ferred to an ice-water bath for no longer than 8 min; this
cycle was repeated 10 times. The cells were dissolved in
196 mL of 50 mM Tris–HCl, pH 7.8 (buffer A). After
incubating the mixture for 30 min on ice, cell debris
was removed by centrifugation (48,000g for 1 h) and
the supernatant containing soluble mtSD was collected.
Purification of recombinant mtSD
All steps of the purification protocol of recombinant
mtSD were performed on ice or at 4 °C. The supernatant
containing soluble mtSD was incubated with 1% (w/v) of
streptomycin sulfate for 30 min and centrifuged at
48,000g for 30 min. The supernatant was dialyzed twice
against buffer A, using a dialysis tubing with molecular
weight exclusion limit of 6000–8000 Da. This sample was
430 I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437
clarified by centrifugation (48,000g for 30 min) and loaded
on a Q-Sepharose fast flow (26 cm · 9.5 cm) column
(Amersham Biosciences) pre-equilibrated with the same
buffer. The column was washed with 5 column volumes
of buffer A and the absorbed material was eluted with a lin-
ear gradient (0–100%) of 20 column volumes of 50 mM
Tris–HCl, pH 7.8, 0.5 M NaCl (buffer B) at 1 mL min
À1
.
The fractions containing mtSD were pooled (55 mL) and
ammonium sulfate was added to a final concentration of
1 M, and clarified by centrifugation (48,000g for 30 min).
The supernatant was loaded on a Phenyl-Sepharose High
Performance (Amersham Biosciences) column pre-equi li-
brated with 50 mM Tris–HCl, pH 7.8, 1 M (NH
4
)
2
SO
4
(buffer C). The column was washed with 5 column volumes
of buffer C and the bound proteins were eluted with a 20-
column volume linear gradient (0–100%) of buffer A at
1 mL min
À1
. The mtSD-containing fractions were pooled
(43 mL), concentrated to less than 4 mL using an Amicon
ultrafiltration cell (MWCO 10,000 Da), and loaded on a
Sephacryl S-200 (26 cm · 60 cm) (Amersham Biosciences)
column pre-equilibrated with buffer A. The recombinant
mtSD protein was eluted in a total volume of 29 mL at a
flow rate of 0.5 mL min
À1
, and loaded on an anion-ex -
change Mono-Q column (Amersham Biosciences) pre-
equilibrated with buffer A. The column was washed with
5 column volumes of buffer A and the absorbed material
was eluted with a 20-col umn volume linear gradient
(0–100%) of buffer B at 1 mL min
À1
. Elution profiles were
followed at 280 and 215 nm. Homogeneous mtSD was elut-
ed in a total volume of 12.5 mL and stored at À20 °C. Pro-
tein purification was monitored by SDS–PAGE [16], and
the protein concentration was determined by the method
of Bradford et al. [17] using the Bio-Rad protein assay
kit (Bio-Rad) and bovine serum albumin as standard.
Enzyme activity assay of mtSD
The screening of fractions of the purification protocol
containing shikimate dehydrogenase was performed by
assaying enzyme activity in the reverse direction in
100 mM Tris–HCl, pH 9.0, at 25 °C. The 500 lL assay
mixture contained 2 mM NADP
+
and 4 mM D-shikimate
(SHK) [18], and the reaction was initiated with addition
of 1 lL of the chromatographic fractions. Measurement
of the NADP
+
-dependent oxidation of SHK to form
NADPH and 3-dehydroshiki mate (DHS) catalyzed by
mtSD was continuously monitored by the increase in
absorbance at 340 nm (e
NADPH
= 6.18 · 10
3
M
À1
cm
À1
).
One unit of enzyme activity (U) is defined as the amount
of enzyme catalyzing the conversion of 1 l mol of NADP
+
per minute at 25 °C.
Determination of apparent kinetic parameters
Determination of the apparent steady-state kinetics
parameters, V
max
and K
m
, for DHS and NADPH in the
forward reaction, was carried out at varying concentration
of one substrate (5, 10, 20, 30, 50, 100, and 200 lM) while
the other was maintained at constant saturation level. The
reverse reaction was performed in the same conditions of
pH (100 mM Tris–HCl buffer, pH 7.0), temperature
(25 °C) and range of concentrations of substrates. The
reaction was initiated with addition of 6 pmol of homoge-
neous M. tuberculosis SD enzyme and was monitored for
1 min. The kinetic data were analyzed by double reciprocal
plots.
Shikimate dehydrogenase activity measurements were
based on decreasing concentration of NADPH upon
DHS reduction or on increasing concentration of NADPH
upon SHK oxidation. The reaction catalyzed by mtSD was
continuously monitored by measuring change in absor-
bance at 340 nm (e
NADPH
= 6.18 · 10
3
M
À1
cm
À1
).
Determination of the energy of activation and thermo
stability of recombinant mtSD
The energy of activation (E
a
) was estimated for recom-
binant mtDS by the following Arrhenius equation:
k ¼ Ae
ÀE
a
=RT
, where k is the rate constant of the react ion
at temperature T (in Kelvin), A is a pre-exp onential factor
(related to collision frequency and a steric factor); E
a
is the
activation energy; R is universal gas constant
(8.3145 J K
À1
mol
À1
)ande
ÀE
a
=RT
is the fraction of molecu-
lar collisions that have energy equal to or greater than the
measured energy of activation (E
a
) of the system at a par-
ticular temperature T. The E
a
value was calculated from
the slope of the linear plot logk versus 1/T (K
À1
) fitting
the data to the following equation: log k =(E
a
/
2.3RT) + log A. Measurements of mtSD enzyme activity
were in triplicates and performed in 100 mM Tris–HCl
buffer, pH 7.0, in the presence of saturating concentrations
of SHK (200 lM) and NADP
+
(100 lM) at the following
temperatures: 15, 20, 25, 30, and 37 °C.
For thermal stability determination, mtSD was incubat ed
at 15, 25, 37, and 55 °C and the remaining enzyme activity
was measured at different times of incubation up to 1 h,
monitoring the mtSD reverse reaction in an assay mixture
containing saturati ng concentrations of substrates (200 lM
SHK and 100 lM NADP
+
) in 100 mM Tris–HCl, pH 7.0,
at 25 °C.
Mass spectrometry analysis
The homogeneity of recombinant protein preparation
was assessed by mass spectrometry (MS), employing some
adaptations made to the syst em described by Chassaigne
and Lobinski [19]. Samples were analyzed on a triple quad-
rupole mass spectrometer, model QUATTRO II, equipped
with a standard electrospray (ESI) probe (Micromass,
Altrinchan), adjusted to c a. 250 lL min
À1
. The source of
temperature (80 °C) and needle voltage (3.6 kV) were main-
tained constant throughout the experimental data
collection, applying a drying gas flow (nitrogen) of
200 L h
À1
and a nebulizer gas flow of 20 L h
À1
. The mass
I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437 431
spectrometer was calibrated with intact horse heart myo-
globin and its typical cone-voltage induced fragments.
The subunit molecular mass of M. tuberculosis SD was
determined by ESI-MS, adjusting the mass spectrometer
to give a peak width at half-height of 1 mass unit, and
the cone sample to skimmer lens voltage controlling the
ion transfer to mass analyzer was set to 38 V. About
50 pmol sample was injected into electrospray transport
solvent. The ESI spectrum was obtained in the multi-chan-
nel acquisition mode, scanning from m/z 500 to 2000 at
scan time of 7 s. The mass spectrometer is equipped with
MassLynx and Transform softwares for data acquisition
and spectra handling.
N-terminal amino acid sequencing
The N-terminal amino acid residues of purified recombi-
nant mtSD were identified by automated Edman degrada-
tion sequencing using a PPSQ 21A gas-phase sequencer
(Shimadzu).
Determination of native mtSD molecular mass
The molecular mass of native mtSD homogenous pro-
tein was estimated by gel-permeation chromatography on
a Superdex 200 HR column (1.0 cm · 30 cm) (Amersham
Biosciences). The column was eluted with 50 mM Tris–
HCl containing 0.2 M NaCl, pH 7.8, at a flow rate of
0.4 mL min
À1
. The eluate was monitored at 215 and
280 nm and the column was calibrated with the following
protein standar ds (Amersham Biosciences): ribonuclease
A (13,700 Da) from bovine pancreas, chymotrypsin ogen
(25,000 Da) from bovine pancreas, ovalbumin
(43,000 Da) from hen egg, and albumin (67,000 Da) from
bovine serum. Blue Dextran 2000 was used to determine
the void volume (V
0
). The K
av
value was calculated for
each protein using the equation (V
e
À V
0
)/(V
t
À V
0
), where
V
e
is the elution volume for the protein and V
t
is the total
bed volume, and K
av
was plotted against the logarithm of
standard molecular weights.
Results and discussion
Expression in E. coli BL21 (DE3) of recombinant
M. tuber culosis shikimate dehydrogenase (mtSD) and the
method of disruption of transformed host cells were as
described elsewhere [15]. Recombinant mtSD was purified
as described under Materials and methods and the samples
of each chromatographic step were analyzed by SDS–
PAGE with Coom assie blue staining and assayed for
enzyme activity in the reverse direction, following the
increase in absorbance at 340 nm due to the NADP
+
-de-
pendent oxidation of
D-shikimate to form NADPH and
3-dehydroshikimate. The recombinant protein was purified
8.5-fold (Table 1) to electrophoretic homogeneity (Fig. 1).
The relative mobility of the polypeptide chain in SDS–
PAGE indicates a homogeneous protein with a subunit
molecular mass value of approximately 27 kDa (Fig. 1).
Even though a large amount of cells were needed for recov-
ery of homogeneous target protein in quantities necessary
for kinetic and structural studies, the freeze–thaw method
[20] was previously shown to be, amongst a num ber of
experimental protocols tested to reduce insoluble protein
production, the method of choice to obtain soluble mtSD
in its active form [15]. Approximately 11 mg of homoge-
neous recombinant mtSD could be obtaine d from 49 g of
E. coli BL21 (DE3) host cells followi ng the purification
protocol presented in Table 1, which required four chroma-
tographic steps to obtain homogeneous mtSD. The recom-
binant protein eluted with approximately 50% of 50 mM
Tris–HCl, pH 7.8, containing 0.5 M NaCl (buffer B), con-
sistent with the theoretical value of 5.11 for the mtSD iso-
electric point. Since the theoretical isoelectric point values
for mtSD and E. coli SD (5.12) are quite similar, the
anion-exchange chromatographic step using Q-Sepharose
Fast Flow resin is unlikely to have separated these
Table 1
Purification of M. tuberculosis shikimate dehydrogenase from E. coli BL21 (DE3) [pET23a(+)::aroE] cells
Purification step Total protein (mg) Total activity (U) Specific activity
a
(U mg
À1
) Purification (fold) Yield (%)
Crude extract 1040.4 475.3 0.46 1.0 100.0
Q-Sepharose 82.5 110.0 1.33 2.9 23.1
Phenyl-Sepharose 24.1 69.2 2.87 6.2 14.6
Sephacryl S-200 17.4 65.3 3.75 8.2 13.7
Mono-Q 10.8 42.1 3.90 8.5 8.9
The results presented are for a typical purification protocol from 49 g of E. coli host cells.
a
UmL
À1
mg.
À1
Fig. 1. SDS–PAGE (12.5%) analysis of pooled fractions from the various
steps of the purification protocol of mtSD. Lane 1, crude extract (50 lg);
lane 2, Q-Sepharose fast flow ion exchange (30 lg); lane 3, Phenyl-
Sepharose hydrophobic interaction (10 lg); lane 4, Sephacryl S-200 gel
permeation (10 lg); lane 5, Mono-Q ion exchange (17 lg); lane 6, MW
marker High Range (Gibco).
432 I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437
enzymes. Nevertheless, this chromatographic step resulted
in 2.9- fold protein purification (Table 1) and removal of
some noticeable contaminants with subunit molecular
weight values larger than 29 kDa (Fig. 1). The recombinant
mtSD protein in 50 mM Tris–HCl, pH 7.8, containing 1 M
(NH
4
)
2
SO
4
adsorbed to a Phenyl-Sepharose High Perfor-
mance column and was eluted with approximately 57% of
50 mM Tris–HCl, pH 7.8, buffer resulting in a 6.2-fold
purification with removal of substantial amount of con-
taminants (Fig. 1). Gel filtration on Sephacryl S-200 resin
was performed to desalt and further purify the recombi-
nant protein. Elution of fractions containing mtSD from
an anion-exchange Mono-Q column with approximately
50% of 50 mM Tris–HCl, pH 7.8, buffer containing 0.5 M
NaCl resulted in homogeneous mtSD, with a protein yield
of approximately 9% (Table 1). It is noteworthy that there
was no need for elution with coenzyme from an affinity col-
umn as described for E. coli shikimate dehydrogenase [21].
Accordingly, the purification protocol here described rep-
resents an efficient and low-cost method to obtain homoge-
neous mtSD. The purification of M. jannaschii shikimate
dehydrogenase by glutathione and Sepharose Q chroma-
tography yielded, after proteolytic removal of the affinity
tag, a protein with five N-terminal residues resulting from
cloning artifact [22]. The strategy followed to obtain homo-
geneous H. influenzae shikimate dehydrogenase provided a
protein with a C-terminal hexahistidine tag [23]. It ha s been
shown that N- and C-terminal hexahistidine tags have a
noticeable negative effect on protein solubility of recombi-
nant proteins expressed in E. coli [24]. In addition,
His-tagged proteins may have different structure [25] or
biological activity [26] as compared to their native form.
The mtSD cloning strategy we described elsewhere [15]
and protein purification we presented here yielded homoge-
neous polypeptide chain with no extra amino acid residues.
The subunit molecular mass of active mtSD was deter-
mined to be 27,076 Da by electrospray ionization mass
spectrometry (ESI-MS), indicating removal of the N-ter-
minal methionine residue (predicted molecular mass:
27,207 Da). The ESI-MS result also revealed a peak at
54,150 Da, indicating that the enzyme could have a
dimeric form. No peak could be detected at the expected
molecular mass for E. coli SD (29,413 Da) and a degree
of purity of 98% could be estimated by ESI-MS, thus
providing evidence for the identity and purity of the
recombinant protein. The first 11 N-terminal amino acid
residues of mtSD were identified as SEGPKKAGVLG
by the Edman degradation method. This result unambig-
uously identifies the homogeneous recombinant protein
as mtSD and confirms removal of the N-terminal methi-
onine. Modification at the N-termini is a common type
of co-/post-translational alteration of proteins synthe-
sized in prokaryotic cells. Methionine aminopeptidase-
catalyzed cleavage of initiator methionine is usually
directed by the penultimate amino acid residues with
the smallest side chain radii of gyration (glycine, alanine,
serine, threonine, proline, valine, and cysteine) [27].
Removal of N-terminal methionine from mtSD polypep-
tide chain conforms to this rule since serine is the penul-
timate amino acid residue.
The enzymatic activity of homogeneous recombinant
mtSD purified was assayed in the reverse direction by con-
tinuously monitoring the increase in absorbance at 340 nm
upon NADP
+
-dependent oxidation of D-shikimate to form
NADPH and 3-dehydroshikimate. The activity of mtSD
was linearly dependent on sample volume added to the
reaction mixture (Fig. 2), thereby showing that the initial
velocity is proportional to total enzyme concentration
and that true initial velocities are being measured. The
M. tuberculosis SD was stable at À20 °C for at least 1 year.
A value of 58,367 Da for the molecular mass of
homogeneous mtSD protein was estimated by gel filtra-
tion (data not shown). This result suggests that mtSD
is a dimer in solution, in agreement with the ESI-MS
results. Shikimate dehydrogenase from M. jannaschii
has recently been shown to be a dimer in solution [22].
Whereas dehydrogenases usually form oligomers, shikim-
ate dehydrogenase is present as a monomer in both
E. coli [18] and H. influenzae [23].
The apparent kinetics parameters obtained are present-
ed in Table 2. The plots fitted to a hyperbolic equation,
indicating that the recombinant enzyme-catalyzed chemical
Fig. 2. Linear dependence of mtSD activity on homogeneous protein
volume. The rates of mtSD enzyme activity were followed in the reverse
reaction by continuously monitoring the increase of NADPH concentra-
tion at 340 nm. Reactions were started by addition of varying volumes of
homogeneous mtSD protein solution.
Table 2
Apparent kinetic parameters for M. tuberculosis shikimate
dehydrogenase
a
Substrates V
max
(U mg
À1
) K
m
(lM) k
cat
(s
À1
) k
cat
/K
m
(M
À1
s
À1
)
DHS 108 ± 5 31 ± 2 49 ± 2 1.6 (±0.1) · 10
6
NADPH 100 ± 5 10 ± 1 45 ± 2 4.5 (±0.5) · 10
6
SHK 18 ± 1 50.18 ± 0.01 8.2 ± 0.5 1.63 (±0.01) · 10
5
NADP
+
12.9 ± 0.7 22 ± 2 5.9 ± 0.3 2.68 (±0.03) · 10
5
a
All constants were measured in Tris–HCl 100 mM (pH 7.0) at 25 °C.
I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437 433
reaction obeys Michaelis–Menten kinetics for all substrates
(Fig. 3). The K
m
and V
max
values for DHS were found to
be, respectively, 31 lM and 108 U mg
À1
; and for NADPH
they were 10 lM and 100 U mg
À1
. The k
cat
for DHS at
saturating NADPH was 49 s
À1
yielding a k
cat
/K
m
of
1.6 · 10
6
M
À1
s
À1
. The K
m
value for DHS is lower than
the value for SD purified from Pisum sativum (340 lM);
but the K
m
value for NADPH is in agreement with the val-
ue of 4.3 lM determined at pH 7.4 [28]. For the reverse
reaction, the K
m
and V
max
values for SHK were found to
be, respectively, 50.18 lM and 18 U mg
À1
; and for NADP
+
they were 22 lM and 12.9 U mg
À1
. The k
cat
for SHK at
saturating NADP
+
was 8.2 s
À1
yielding a k
cat
/K
m
of
1.63 · 10
5
M
À1
s
À1
. For the reverse reaction catal yzed by
P. sativum SD, the K
m
values were 10.3 lM for NADP
+
and 600 lM for SHK [28]. The K
m
value for SHK is thus
10 times larger for P. sativum SD than for mtSD. A com-
parison of the SHK and NADP
+
K
m
values for E. coli
SD [29] with those obtained for mtSD shows that these
two en zymes catalyze the reverse reaction with similar
kinetic constants. A value of 19.6 for the apparent
equilibrium constant (K
eq
) under the experimental
conditions given in the Materials and methods section
was estimated by the Haldane equation using the apparent
steady-state kinetic parameters for mtSD. An apparent
equilibrium constant value of 10.3 for the reaction at pH
7.4 has been determined by finding a mixture of substrate
concentrations that showed no measurable chan ge in
optical density at 340 nm [28]. However, it should be
pointed out that the Haldane equation we used here is
the kinetic relationship for rapid-equilibrium random Bi
Bi system, and it was implicitly assumed that the dissocia-
tion constant value for a substr ate binding second is not
changed by the substrate binding first [30].
The temperature effects on recombinant homogeneous
mtSD stability are presented in Fig. 5. Aliquots were removed
for assay of residual enzyme activity after heating for 1, 5, 10,
15, 30, 45, and 60 min at temperature values of 15, 25, 37, and
55 °C. The homogeneous mtSD is very stable, maintaining
the specific activity, measured at 25 °C, unchanged in the
reverse reaction up to 37 °C for at least 1 h of incubation
(Fig. 4). However, at 55 °C, there is a gradual loss of recombi-
nant mtSD biological activity, with only 20% of the initial
enzyme activity remaining after 1 h of incubation.
Fig. 3. Reciprocal plots with 3-dehydroshikimate (DHS), NADP
+
, D-shikimate (SHK) or NADPH as variable substrate (5, 10, 20, 30, 50, 100, and
200 lM) while the concentration of the other substrate was maintained at constant saturation level in the forward and reverse reaction. The enzyme
activity was assayed at 25 °C in 100 mM Tris–HCl buffer, pH 7.0. The reaction catalyzed by mtSD was continuously monitored by the absorbance at
340 nm (e
NADPH
= 6.18 · 10
3
M
À1
cm
À1
). The kinetics data were analyzed by linear regression fit Michaelis–Menten kinetics. (A) 3-dehydroshikimate as
variable substrate (5, 10, 20, 30, 50, 100, and 200 lM) while the NADPH was maintained at constant saturation level (200 lM) in the forward reaction. (B)
NADPH as variable substrate (5, 10, 20, 30, 50, 100, and 200 lM) while 3-dehydroshikimate was maintained at constant saturation level (200 lM) in the
forward reaction. (C) Shikimate as variable substrate (5, 10, 20, 30, 50, 100, and 200 lM) while NADP
+
was maintained at constant saturation level
(100 lM) in the forward reaction. (D) NADP
+
as variable substrate (5, 10, 20, 30, 50, 100, and 200 lM) while shikimate was maintained at constant near-
saturation level (100 lM) in the forward reaction.
434 I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437
From the linear plo ts of logk versus 1/T (K
À1
), a val-
ue of 35.2 kJ mol
À1
for E
a
was obtained for mtSD
(Fig. 5). It should be pointed out that the E
a
value cal-
culated from the Arrhenius plot is an apparent or ‘‘aver-
age value,’’ and that the pre-exponential factor (A) was
considered as temperature-independent in the tempera-
ture range used in our experiments. Keeping that in
mind, 35.2 kJ mol
À1
can be considered as the minimal
amount of energy required to initiate the mtSD-catalyzed
chemical reaction, since we measured the enzyme activity
at saturating concentrations of substrates (200 lM SHK
and 100 lM NADP
+
). Interestingly, the Arrhenius plot
is linear indicating no change in the rate-limiting step
of the mtSD-catalyzed chemical reaction at different tem-
peratures. In addition, there was no su dden drop in the
Arrhenius plot at low 1/T (high T) va lues that could
indicate protein denaturation, which is consistent with
mtSD thermal stability up to 55 °C.
The three-dimensional structures of shikimate dehydro-
genase from E. coli [29,31], H. influenzae [23,32],and
M. jannaschii [22] have been determined. Sequence align-
ment of shikimate dehydrogenas e from M. tuberculosis
H37Rv strain, E. coli , H. influenzae, M. jannaschii was car-
ried out using ClustalW [33,34]. The identity between the
SD sequences from M. tuberculosis and E. coli is 24%. The
substrate-binding site in the E. coli SD has been identified
by the position of the nicotinamide ring of the cofactor and
was delineated almost entirely by residues from the N-termi-
nal domain [29]. This binding site is in a pocket where most of
the residues absolutely conserved in the shikimate dehydro-
genase (SDH) family are located, i.e., Ser14, Ser16, Lys65,
Asn86, Thr101, Asp102, and Gln244 (E. coli SD numbering).
These residues are conserved in the mtSD polypeptide
sequence (Fig. 6) corresponding to the following residues:
Ser18, Ser20, Lys69, Asn90, Thr104, Asp105, and Gln243
(M. tuberculosis SD numbering). How ever, the 3-dehydros-
hikimate-binding site has not unambig uously been shown
for E. coli shikimate dehydrogenase since it was inferred
from DTT and sulfate ions bound to the crystal and position
of the nicotinamide ring of NADP
+
[29]. The sequence
identity (24%) between mtSD and M. jannaschii SD is similar
to the identity betw een the sequences of M. tuberculosis and
E. coli SDs. The N-terminal domain (domain I) of M. jann-
aschii SD is responsible for 3-dehydroshikimate substrate
binding [22]. The putative active site residues within domain
I that are likely to be involved with substrate binding are
invariant polar residues include Lys70, Asn91, and Asp106
(M. jannaschii SD numbering). Further examination of the
active site cleft of M. jannaschii SD revealed another strictly
conserved residue, Gln254. These residues correspond to
Lys69, Asn90, Asp105, and Gln244 in M. tuberculosis SD.
It has been proposed that these residues are likely to be
involved in catalytic reduction of DHS to SHK catalyzed
by M. jannaschii SD [22]. The amino acids Ser14, Ser16,
Lys65, Asn85, Asp102, and Gln245 (H. influenzae SD
numbering) have been suggested as some of the potential
residues involved in 3-dehydroshikimate-binding site in
H. influenzae SD [23]. All these residues are conserved in
the mtSD polypeptide sequence (Fig. 6), and corres pond to
the same residues observed in the E. coli SD sequence. More
recently, the crystal structure of a newly characterized
shikimate dehydrogenase-like protein (HI0607) from
H. influenzae has been determined [32], and the conserved
residues Lys67, Asn88, Asp103, and Gln242 have be en
proposed to be involved in either catalysis or substrate bind-
ing. It should be pointed out that the three-dimensional
structures of shikimate dehydrogenases determined were
in complex with the cofactor and no three-dimensional
structure in complex with 3-dehydroshikimate/shikimate
molecules have been reported to date.
Fig. 5. The activation energy (E
a
) estimated for homogeneous
M. tuberculosis SD enzyme by the Arrhenius equation. All data were
obtained in the reverse reaction in 100 mM Tris–HCl, pH 7.0, containing
200 lM SHK, 100 lM NADP
+
, and 0.24 nM of homogeneous
M. tuberculosis SD enzyme and the reaction was monitored for 1 min at
the following temperatures: 288, 293, 298, 303, and 310 K.
Fig. 4. Heat stability of mtSD activity as a function of time (minutes) of
incubation at different temperatures. Incubation temperatures were as
follows: 15 °C(d), 25 °C(m), 37 °C(.), and 55 °C(n). All reactions were
carried out in 100 mM Tris–HCl, pH 7.0, assay mixture containing
200 lM SHK, 100 lM NADP
+
, and 0.24 nM of homogeneous
M. tuberculosis SD enzyme and the reaction was measured for 1 min at
25 °C.
I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437 435
The work presented here describes, to the best of our
knowledge, the first purification protocol of recombinant
M. tuberculosis SD to homogeneity and determ ination of
its oligomeric state. In addition, here we present a detailed
characterization of the homogeneous mtSD. It should be
pointed out that expression of mtSD in soluble and active
form proved to be laborious to achieve [15]. Protein pro-
duction and crystallization must be optimized if structural
genomics will ever reach its goal of solving the three-di-
mensional structure of the whole proteome encoded by a
given genome [35]. Unfortunately, even when a genome
can be sequenced, only up to 20% of the protein targets
can produce soluble proteins under very basic experimental
conditions [36]. Thus, expression of proteins in soluble
form has been identified as an important bottleneck in
efforts to determine biological activity and crystal structure
of M. tuberculosis proteins [37].
The protocol for mtSD purification presented here
should provide protein in qua ntities necessary for further
enzymological studies and three-dimensional structure
determination efforts. Dete rmination of mtSD kinetic
mechanism by steady-state and pre-steady-state kinetics,
and isotope effects, site-directed mutagenesis, and chemical
rescue will allow elucidation of its chemical and catalytic
mechanism. The three-dimensional structures of shikimate
dehydrogenases from E. coli [29,31], H. influenzae [23,32],
and M. jannaschii [22] should facilitate screening of exper-
imental con ditions to obtain crystals of mtSD protein and
may provide templates for mtSD structure determination
by molecular replac ement. All of the enzymes that make
up the shikimate pathway are potential targets for the
design of novel drugs directed against pathogenic bacteria.
The struc ture of M. tuberculosis shikimate dehydrogenase
and the understanding of the enzyme mode of action will
be used as a platform for the design of effective inhibitors
of this pathway aiming at the development of antitubercu-
lar agents.
Acknowledgments
Financial support for this work was provide by Millen-
nium Initiative Program MCT-CNPq, Ministry of Health-
Department of Science and Technology-UNESCO (Brazil)
to D.S.S. and L.A.B. D.S.S. and L.A.B. also acknowledge
grants awarded by PRONEX, CNPq/FAPERGS. D.S.S.
(CNPq, 304051/1975-06), L.A.B. (CNPq, 520182/99-5),
and M.S.P. (CNPq, 500079/90 -0) are researchers awardees
from the National Research Council (CNPq) of Brazil. We
Fig. 6. Multiple sequence alignment of SD from M. tuberculosis using ClustalW. The boxes denote residues conserved in the SDH family involved in
catalytic reduction of DHS to SHK. This sequence alignment was constructed using the following sequences from GenBank: E. coli K12 AroE
(NP_417740, residues 1–172), H. influenzae AroE (YP_248346, residues 1–268), M. jannaschii AroE (NP_248077, residues 1–282), and M. tuberculosis
AroE (CAB06186, residues 1–269). The symbols are as follows: * = identity, : = strong similarity, . = week similarity. These nomenclature is based on the
properties of aminoacids using the substitution matrix BLOSUM [34].
436 I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437
thank Professor John W. Frost, Department of Chemistry
of Michigan State University, for his generous gift of
3-dehydroshikimate substrate.
References
[1] World Health Organization, World Health Report on Infections
Diseases: Removing Obstacles to Healthy Development. Atar,
Switzerland. WHO/CDS/99.1, 1999.
[2] C. Dye, S. Scheele, P. Dolin, V. Pathania, M.C. Raviglione, Global
burden of tuberculosis: estimated incidence, prevalence, and mortality
by country, JAMA 282 (1999) 677–686.
[3] B.R. Bloom, C.J.L. Murray, Tuberculosis: commentary on a
reemergent killer, Science 257 (1998) 1055–1064.
[4] A. Knigge, H. Morr, A. Kilian, Tuberculosis and HIV/AIDS epidem-
ics: opportunities for cross support, Deutsche Gesellschaft Fu
¨
r Tech-
nische Zusammenarbeit (GTZ) GmbH (2000) 1–7 (<www.gtz.de/>).
[5] L.A. Basso, J.S. Blanchard, Resistance to antitubercular drugs, Adv.
Exp. Med. Biol. 456 (1998) 115–144.
[6] J.D. McKinney Jr., W.R. Jacobs, B.R. Bloom, Persisting problems in
tuberculosis, in: R.M. Krause (Ed.), Emerging Infections, Academic
Press, New York, 1998, pp. 51–146.
[7] World Health Organization, Anti-tuberculosis drug resistance in the
world: Third Global Report (2004).
[8] K. Duncan, Progress in TB drug development and what is still
needed, Tuberculosis 83 (2003) 201–207.
[9] R.J. OÕBrien, P.P. Nunn, The need for new drugs against tuberculosis,
Am. J. Respir. Crit. Care Med. 162 (2001) 1055–1058.
[10] R. Bentley, The shikimate pathway—a metabolic tree with many
branches, Crit. Rev. Biochem. Mol. Biol. 25 (1990) 307–384.
[11] C. Ratledge, Nutrition, growth and metabolism, in: C. Ratledge, J.L.
Stanford (Eds.), The Biology of the Mycobacteria, vol. 1, Academic
Press, London, 1982, pp. 185–271.
[12] J.J. de Voos, K. Rutter, B.G. Schroder, H. Su, Y. Zhu, C.E. Barry III,
The salicylate-derived mycobactin siderophores of Mycobacterium
tuberculosis are essential for growth in macrophages, Proc. Natl.
Acad. Sci. USA 97 (2000) 1252–1257.
[13] T. Parish, N.G. Stoker, The common aromatic amino acid biosyn-
thesis pathway is essential in Mycobacterium tuberculosis, Microbiol-
ogy 148 (2002) 3069–3077.
[14] S.T. Cole, R. Brosch, J. Parkhill, T. Garnier, C. Churcher, D. Harris,
S.V. Gordon, K. Eiglmeier, S. Gas, C.E. Barry Jr., F. Tekaia, K.
Badcock, D. Basham, D. Brown, T. Chillingworth, R. Connor, R.
Davies, K. Devlin, T. Feltwell, S. Gentles, N. Hamlin, S. Holroyd, T.
Hornsby, K. Jagels, B.G. Barrell, Deciphering the biology of
Mycobacterium tuberculosis from the complete genome sequence,
Nature 393 (1998) 537–544.
[15] M.L. Magalha
˜
es, C.P. Pereira, L.A. Basso, D.S. Santos, Cloning and
overexpression of functional shikimate dehydrogenase (EC 1.1.1.25)
from Mycobacterium tuberculosis H37Rv, Protein Expr. Purif. 26
(2002) 59–64.
[16] U.K. Laemmli, Cleavage of structural proteins during the assembly of
the head of bacteriophage T4, Nature 227 (1970) 680–685.
[17] M.M. Bradford, R.A. McRorie, W.L. Williams, A rapid and sensitive
method for the quantitation of microgram quantities of protein
utilizing the principle of protein-dye binding, Anal. Biochem. 72
(1976) 248–254.
[18] S. Chaudhuri, J.R. Coggins, The purification of shikimate
dehydrogenase from Escherichia coli, Biochem. J. 226 (1985)
217–223.
[19] H. Chassaigne, R. Lobinski, Characterization of horse kidney
metallothionein isoforms by electrospray MS and reversed-phase
HPLC–electrospray MS, Analyst 123 (1998) 2125–2130.
[20] B.H. Johnson, M.H. Hecht, Recombinant proteins can be isolated
from E. coli cells by repeated cycles of freezing and thawing, Bio/
Technology 12 (1994) 1357–1360.
[21] J. Maclean, S.A. Campbell, K. Pollock, S. Chackrewarthy, J.R.
Coggins, A.J. Lapthorn, Crystallization and preliminary X-ray
analysis of shikimate dehydrogenase from Escherichia coli, Acta
Crystallogr. Sect. D. Biol. Crystallogr. 56 (2000) 512–515.
[22] A.K. Padyana, S.K. Burley, Crystal structure of shikimate 5-
dehydrogenase (SDH) bound to NADP: insights into function and
evolution, Structure 11 (2003) 1005–1013.
[23] S. Ye, F.v. Delft, A. Brooun, M.W. Knuth, R.V. Swanson, D.E.
McRee, The crystal structure of shikimate dehydrogenase (AroE)
reveals a unique NADPH binding mode, J. Bacteriol. 185 (2003)
4144–4151.
[24] E.A. Woestenenk, M. Hammarstro
¨
m, S. van dn Berg, T. Ha
¨
rd, H.
Berglund, His tag effect on solubility of human proteins produced in
Escherichia coli: a comparison between four expression vectors, J.
Struct. Funct. Genomics 5 (2004) 217–229.
[25] A. Chant, C.M. Kraemer-Pecore, R. Watkin, G.G. Kneale, Attach-
ment of a histidine tag to the minimal zinc finger protein of the
Aspergillus nidulans gene regulatory protein AreA causes a confor-
mational change at the DNA-binding site, Protein Expr. Purif. 39
(2005) 152–159.
[26] I. Fonda, M. Kenig, V. Gaberck-Porekar, P. Prostovaek, V. Menart,
Attachment of histidine tags to recombinant tumor necrosis factor-
alphadrasticallychangesitsproperties,Sci.WorldJ. 2(2002)1312–1325.
[27] P.-H. Hirel, J.-M. Schmitter, P. Dessen, G. Fayat, S. Blanquet, Extent
of N-terminal methionine excision from
Escherichia coli proteins is
governed by the side-chain length of the penultimate amino acid,
Proc. Natl. Acad. Sci. USA 86 (1989) 8247–8251.
[28] D. Balinsky, A.W. Dennis, W.W. Cleland, Kinetic and isotope-
exchange studies on shikimate dehydrogenase from Pisum sativum,
Biochemistry 10 (1971) 1947–1952.
[29] G. Michel, A.W. Roszak, V. Sauve
´
, J. Maclean, A. Matte, J.R.
Coggins, M. Cygler, A.J. Lapthorn, Structures of shikimate dehy-
drogenase AroE and its paralog YdiB, J. Biol. Chem. 278 (2003)
19463–19472.
[30] W.W. Cleland, An analysis of Haldane relationships, Methods
Enzymol. 87 (1982) 366–369.
[31] J. Benach, I. Lee, W. Edstron, A.P. Kuzin, Y. Chiang, T.B. Acton,
G.T. Montelione, J.F. Hunt, The 2.3 A
˚
crystal structure of the
shikimate 5-dehydrogenase orthologue YdiB from Escherichia coli
suggests a novel catalytic environment for na NAD-dependent
dehydrogenase, J. Biol. Chem. 278 (2003) 19176–19182.
[32] S. Singh, S. Korolev, O. Koroleva, T. Zarembinski, F. Collart, A.
Joachimiak, D. Christendat, Crystal structure of a novel shikimate
dehydrogenase from Haemophilus influenzae, J. Biol. Chem. 280
(2005) 17101–17108.
[33] BLAST: S.F. Altschul, T.L. Madden, A.A. Scha
¨
ffer, J. Zhang, Z.
Zhang, W. Miller, D.J. Lipman, Gapped BLAST and PSIBLAST: a
new generation of protein database search programs, Nucleic Acids
Res. 25 (1997) 3389–3402.
[34] CLUSTALW: J.D. Thompson, D.G. Higgins, T.J. Gibson, CLUS-
TAL W: improving the sensitivity of progressive multiple sequence
alignment through sequence weighting, position-specific gap penalties
and weight matrix choice, Nucleic Acids Res. 22 (1994) 4673–4680.
[35] W.R.A. Taylor, A Ôperiodic tableÕ for protein structures, Nature 415
(2002) 657–660.
[36] S.A. Lesley, P. Kuhn, A. Godzik, A.M. Deacon, I. Mathews, A.
Kreusch, G. Spraggon, H.E. Klock, D. McMullan, T. Shin, J.
Vincent, A. Robb, L.S. Brinen, M.D. Miller, T.M. McPhillips, M.A.
Miller, D. Scheibe, J.M. Canaves, C. Guda, L. Jaroszewski, T.L.
Selby, M.A. Elsliger, J. Wooley, S.S. Taylor, K.O. Hodgson, I.A.
Wilson, P.G. Schultz, R.C. Stevens, Structural genomics of the
Thermotoga maritima proteome implemented in a high-throughput
structure determination pipeline, Proc. Natl. Acad. Sci. USA 99
(2002) 11664–11669.
[37] R. Vicentelli, C. Bignon, A. Gruez, S. Canaan, G. Sulzenbacher, M.
Tegoni, V. Campanacci, C. Cambillau, Medium-scale structural
genomics: strategies for protein expression and crystallization, Acc.
Chem. Res. 36 (2003) 165–172.
I.O. Fonseca et al. / Protein Expression and Purification 46 (2006) 429–437 437
Capítulo II
Este artigo será submetido para a revista ARCHIVES OF
BIOCHEMISTRY AND BIOPHYSICS
Fonseca I.O., Silva R.G., Fernandes C., Souza O.N., Basso L.A.,
Santos D.S. “Shikimate Dehydrogenase from Mycobacterium
tuberculosis H37Rv: Kinetic and Chemical Mechanisms.
23
Kinetic and chemical mechanisms of shikimate dehydrogenase from
Mycobacterium tuberculosis
Isabel O. Fonseca
π
, Rafael G. Silva
π
, Claudia Fernandes
, Osmar Norberto de Souza
, Luiz A.
Basso
‡*
, and Diógenes S. Santos
‡*
Centro de Pequisas em Biologia Molecular e Funcional, Pontifícia Universidade Católica do Rio
Grande do Sul (PUCRS), Porto Alegre - RS 90619-900, Brazil.
π
Programa de Pós-Graduação em Ciências Biológicas: Bioquímica, Universidade Federal do Rio
Grande do Sul (UFRGS).
Laboratório de Bioinformática, Modelagem e Simulação de Biossistemas – LABIO, Faculdade
de Informática, PUCRS, Porto Alegre - RS 90619-900, Brazil.
Short Title: M. tuberculosis shikimate dehydrogenase mode of action
1
Address correspondence to: Luiz A. Basso or Diógenes S. Santos, Centro de Pesquisas
em Biologia Molecular e Funcional, Pontifícia Universidade Católica do Rio Grande do Sul
(PUCRS), Porto Alegre - RS - 90619-900, Brazil. Phone/Fax: +55 51 33203629; E-mail:
ABSTRACT
Mycobacterium tuberculosis shikimate dehydrogenase (MtbSD) catalyzes the fourth
reaction in the shikimate pathway, the NADPH-dependent reduction of 3-
dehydroshikimate. To gather information on the kinetic mechanism, initial velocity
patterns, product inhibition, and primary deuterium kinetic isotope effect studies were
performed and the results suggested a steady-state ordered bi-bi kinetic mechanism. The
magnitudes of both primary and solvent kinetic isotope effects indicated that the hydride
transferred from NADPH and protons transferred from the solvent in the catalytic cycle are
not significantly rate limiting in the overall reaction. Proton inventory analysis indicates
that one proton gives rise to solvent isotope effects. Multiple isotope effect studies indicate
that both hydride and proton transfers are concerted. The pH profiles revealed that
acid/base chemistry takes place in catalysis and substrate binding. The MtbSD 3D model
was obtained in silico by homology modeling. Kinetic and chemical mechanisms for
MtbSD are proposed on the basis of experimental data.
Keywords: shikimate dehydrogenase, tuberculosis, drug target, 3-dehydroshikimate,
shikimate, kinetic mechanism, chemical mechanism, Mycobacterium tuberculosis.
2
Introduction
Tuberculosis
1
(TB) is a pandemic, which even today, remains a major global health
concern. Its causative agent, Mycobacterium tuberculosis, is one of the most prolific
infectious agents affecting humans. A third of the world’s population is thought to host M.
tuberculosis and approximately 30 million people have died from the disease in the past
decade [1]. The treatment of multidrug-resistant TB (MDR-TB), defined as resistant to at
least isoniazid and rifampicin [2], 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 chemotherapy regimens [3]. 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 2% were XDR, and that XDR-TB has a wide geographic
distribution [4]. New antimycobacterial agents are thus needed to improve the treatment of
MDR- and XDR-TB, and to provide more effective treatment of latent tuberculosis
infection [5].
The availability of the complete genome sequence of M. tuberculosis allowed the
identification of molecular targets against which new therapeutical agents may be
developed [6]. Homologues to the seven enzymes of the shikimate pathway were identified
in the genome sequence of M. tuberculosis. This pathway has been shown to be essential
for the viability of M. tuberculosis [7]. Accordingly, the essentiality of the mycobacterial
shikimate pathway and its absence from human host indicate that any enzyme of this
pathway represents a promising target for the development of non-toxic antimycobacterial
3
agents. However, an understanding of the kinetic mechanism and availability of structural
models are needed to help the rational inhibitor design to progress more rapidly [8].
Shikimate dehydrogenase (EC 1.1.1.25), the fourth enzyme in the shikimate
biosynthesis pathway, catalyzes the NADPH-dependent reduction of 3-dehydroshikimate
(DHS) to shikimate (SHK) (Fig. 1). We have previously reported cloning and expression of
M. tuberculosis SD (MtbSD) [9]. In addition, measurements of the NADP
+
-dependent
oxidation of shikimate to 3-dehydroshikimate catalyzed by recombinant MtbSD confirmed
the correct assignment to the structural gene encoding SD in M. tuberculosis [9]. More
recently, we have reported purification to homogeneity of recombinant functional MtbSD,
N-terminal amino acid sequencing, electrospray ionization mass spectrometry analysis,
chromatographic size exclusion, and enzyme activity measurements that unambiguously
demonstrated the identity of recombinant MtbSD, its oligomeric state, the apparent kinetic
parameters for all substrates, the thermal stability, and the activation energy for the
enzyme-catalyzed chemical reaction [10].
In the present work, initial velocity patterns in the forward direction, product
inhibition studies, primary deuterium kinetic isotope effects, solvent kinetic isotope effects,
proton inventory, multiple kinetic isotope effects, pH rate profiles, and homology modeling
were performed to investigate the MtbSD kinetic and chemical mechanisms. The data here
presented provide a framework on which to base the rational design of inhibitors with
possible antitubercular activity.
4
Materials and Methods
Enzymatic Assay for MtbSD - All enzyme activity assays were carried out at 25°C in 100
mM potassium phosphate (pH 7.3) by monitoring the decrease in absorbance at 340 nm (ε
= 6220 M
-1
cm
-1
) accompanying the conversion of NADPH to NADP
+
in the presence of
DHS and MtbSD. The final enzyme concentration was 1.8 nM. To obtain initial velocity
measurements, rate values were determined under experimental conditions in which less
than 5% of substrate was consumed.
Initial Velocity and Product Inhibition - To determine the steady-state kinetic parameters
and initial velocity patterns, MtbSD activity was measured in the presence of varying
concentrations of one substrate and several fixed-varied concentrations of the other.
Product inhibition patterns were determined by measuring initial rates at five
concentrations of one substrate, fixed non-saturating concentration of the co-substrate, and
fixed-varying levels of product (either NADP
+
or SHK).
Kinetic Isotope Effects and Proton Inventory - The synthesis of [4S-
2
H]NADPH was
performed as described by Ottolina et al. (1989) [11]. Both NADPH and NADPD
substrates were purified on an FPLC Mono-Q column (GE Healthcare) as previously
described, and the fractions with absorbance ratios A
260nm
/A
340nm
2.3 were pooled [12].
All measurements were performed in duplicate. Primary deuterium kinetic isotope effects
were determined by measuring initial rates in the presence of varying concentrations of one
substrate and five fixed concentrations of the other, with either NADPH or NADPD.
Solvent kinetic isotope effects were determined by measuring initial velocities using a
5
saturating level of one substrate and varying concentrations of the other in either H
2
O or 90
atom % D
2
O. Multiple kinetic isotope effects were obtained by determining the primary
isotope effects using D
2
O as solvent. The proton inventory was determined using saturating
concentrations of DHS and NADPH at various mole fractions of D
2
O. Each proton
inventory individual initial rate datum is the average of triplicate measurements.
pH-Rate Profiles - To determinate the pH dependence of k
cat
and k
cat
/K
m
, initial velocities
were measured in the presence of varying concentrations of one substrate and a saturating
level of the other at different pHs in a mixture of potassium phosphate and boric acid. The
mixed buffers were used over the following pH values: 5.5, 6.0, 6.5, 7.0, 7.5, 8.0, 8.5, 9.0,
9.5, and 10.0. The MtbSD enzyme had been incubated in buffers at all the pH values
mentioned above and assayed under standard conditions to ensure enzyme stability at the
tested pHs.
Data Analysis - Initial velocity kinetic data were fitted to appropriate equations by using the
nonlinear regression function of SigmaPlot 2000 (SPSS, Inc.). Substrate hyperbolic
saturation curves at a single concentration of the fixed substrate and varying concentrations
of the other substrate were fitted to equation 1. Intersecting initial velocity patterns were
fitted to equation 2, which describes a sequential mechanism. For equations 1 and 2, v is the
measured reaction velocity, V is the maximal velocity, A and B are the concentrations of
substrates (DHS and NADPH), K
a
, and K
b
are the corresponding Michaelis- Menten
constants, and K
ia
is the dissociation constant for substrate A.
v = VA / (K
a
+ A) (Eq. 1)
v = VAB / (K
a
B + K
b
A + K
ia
K
b
+ AB) (Eq. 2)
6
Competitive and noncompetitive inhibition data were fitted to, respectively,
equation 3 and 4. For equations 3 and 4, I represent the inhibitor concentration, and K
is
and
K
ii
are the slope and intercept inhibition constants, respectively.
v = VA / [K
a
(1 + I/K
is
) + A] (Eq. 3)
v = VA / [K
a
(1 + I/K
is
) + A (1 + I/K
ii
)] (Eq. 4)
The kinetic isotope effect data were fitted to equations 5 or 6, which describe,
respectively, effects on both V and V/K and on V only, where Fi represents the fraction of
isotopic label, and E
VK
is the isotope effect on V/K minus one and E
V
is the isotope effect on
V minus one.
v = VA / [K (1 + FiE
V/K
) + A (1 + FiE
V
)] (Eq. 5)
v = VA / [K + A (1 + FiE
V
)] (Eq. 6)
Data for pH profiles data that showed a decrease in logk
cat
and logk
cat
/K
m
with a
slope of -1 as the pH values increased were fitted to equation 7, where y is the apparent
kinetic parameter, C is the pH-independent plateau value of y, H is the hydrogen ion
concentration, and K
b
is the apparent base dissociation constant for ionizing groups.
log y = log [C/(1 + K
b
/H)] (Eq. 7)
Homology Modeling - Primarily, homology modeling is the identification and selection of
template proteins from the Protein Data Bank (PDB) [13] that are related to the target
sequence. The program Blastp [14] was used to search for templates. Multiple sequence
alignment comparisons were carried out to improve the sensitivity of the search and to find
regions with high similarity using ClustalW [15]. The program MODELLER6v2 [16] was
used to build the protein models, using the standard protocol of the comparative protein
structure modeling methodology [17]. The best model of each enzyme was evaluated and
7
selected according to their stereochemical quality analyzed with PROCHECK [18].
Validation of the 3D profiles of the models was performed with VERIFY 3D [19]. Here,
the structure prediction of MtbSD was based on 3D structure for the homologous SD from
Escherichia coli (E. coli) (PDB ID: 1NYT), experimentally determined by X-ray
diffraction at 1.5Å resolution [20]. Root-mean square deviation (RMSD) between template
and model structures was calculated using SwissPdbViewer [21]. Drawing of protein
structures were performed with the program SwissPdbViewer [21].
8
Results
Initial Velocity and Product Inhibition Patterns - The constants for steady-state
kinetics of MtbSD were determined under initial velocity experimental conditions. When
DHS was varied at fixed-varied NADPH concentrations, the lines intersected on the x-axis
and on the left of the y-axis (Fig. 2A); when NADPH was varied with fixed-varied DHS
concentrations, the lines intersected below the x-axis and on the left of the y-axis (Fig. 2B).
The data were fitted to equation 2, and yielded the following values: K
DHS
= 44 ± 3
μ
M,
K
iaDHS
= 30 ± 9, K
NADPH
= 34 ± 2
μ
M, k
cat
= 78 ± 2 s
-1
, k
cat
/K
DHS
= 1.8 x 10
6
M
-1
s
-1
, and
k
cat
/K
NADPH
= 2.3 x 10
6
M
-1
s
-1
.
Product inhibition experiments were performed and the data fitted to equation 3 or
4. The results are given in Table 1. NADP
+
is a competitive inhibitor (C) and a
noncompetitive inhibitor (NC) versus NADPH and DHS, respectively. For SHK product,
the inhibition pattern is C and NC versus, respectively, DHS and NADPH.
Kinetic Isotope Effects and Proton Inventory - The primary deuterium kinetic
isotope effects on V/K and V for both substrates are presented in Table 2. The primary
kinetic isotope effect values of 1.8 for
D
V
DHS
and 1.5 for
D
V
NADPH
using [4S-
2
H]NADPH
indicate that the C
4
-proS hydride (B side) is transferred to DHS in the oxy-reduction
reaction catalyzed by MtbSD.
The
D
(V/K
app
)
DHS
values decreased to a limiting value of 1.0 ± 0.03 as the
concentrations of either NADPH or NADPD increased (Fig. 3), whereas
D
(V/K
app
)
NADPH
(1.4 ± 0.3) was independent of fixed-varied concentrations of DHS (Fig. 3 - inset).
9
Solvent kinetic deuterium isotope effects were determined at pH 7.3 (Fig. 4A and
Fig. 4B), a pH region in which the kinetic parameters are independent of small changes in
pH. Small solvent kinetic isotope effects on V were observed for DHS (1.5 ± 0.3) and
NADPH (1.3 ± 0.2), whereas no solvent effects on V/K were observed (Table 2). The
proton inventory, a relationship between V and the mole fraction of D
2
O, was linear for
MtbSD (Fig. 4A - inset).
In order to distinguish between stepwise and concerted mechanisms, multiple
isotope effects were evaluated measuring the primary isotope effects for both substrates in
either D
2
O (Fig. 5A and 5B) or H
2
O (Fig. 5A and 5B - insets), and the results are
summarized in Table 2. The
D
V/K
NADPH
value in H
2
O (1.4 ± 0.3) was within standard error
of the
D
V/K
NADPH
value in D
2
O (1.5 ± 0.1), whereas
D
V
NADPH
in D
2
O (2.5) increased as
compared to the value obtained in H
2
O (1.5). The magnitude of the
D
V
DHS
value in D
2
O
(2.6) also increased, and, most important, running the reaction in D
2
O allowed the
observation of a small, but significant,
D
V/K
DHS
, (1.3 ± 0.1) in contrast to the unity obtained
in H
2
O.
pH Rate Profiles - To probe the role of acid/base chemistry in the mechanism of
MtbSD, the pH dependence of k
cat
and k
cat
/K
m
for DHS and NADPH was determined over
the pH range of 5.5 pH 10.0. The pH profiles of k
cat
for both substrates are very similar
and decrease at high pHs with a slope of -1 (Fig. 6A), demonstrating that deprotonation of a
single residue with apparent pK
a
value of 8.9 ± 0.1 abolishes MtbSD catalytic activity. The
k
cat
/K
m
profiles for DHS and NADPH show a decrease at high pH values with a slope of -1
10
(Fig. 6B and Fig. 6C), suggesting that deprotonation of a single ionizable group with pK
a
value of 9.1 ± 0.1 diminishes DHS and NADPH binding.
Homology Modeling -In the search for templates to perform the homology modeling
we found three candidate structures, namely, E. coli SD (PDB ID: 1NYT) [20], E. coli
YdiB (PDB ID: 109B) [20] and Methanococcus jannaschii SD (PDB ID: 1NVT) [22]. The
Blastp results listed M. jannaschii SD (PDB ID: 1NVT) as the best template for MtbSD.
However, the structure coordinates were not available at the time of this modeling work.
The second best result was YdiB (PDB ID: 109B), although its function is different from
that of MtbSD. YdiB was characterized as a dual specificity quinate/shikimate
dehydrogenase that utilizes either NAD
+
or NADP
+
as cofactor. Multiple sequence
alignment, with the insertion of two gap regions into the MtbSD sequence, illustrates this
difference (Fig. 7). Segments of the target sequence which have no equivalent in the
template are the most difficult regions to model. Hence, the E. coli SD (PDB ID: 1NYT),
solved experimentally by X-ray diffraction at 1.5Å resolution [20], was used as a template
to model the 3D structure of MtbSD. They both have the same function and a sequence
identity of approximately 25%, the limit usually allowed for comparative protein structure
modeling [14].
Ten models of the enzyme were built, and evaluated by PROCHECK [18] and
VERIFY 3D [19] to choose the best one. MtbSD contains 269 amino acid residues. Out of
220 non-glycine and non-proline residues, 199 or 90.5% were located in the most favored
regions of the Ramachandran plot. The best model of MtbSD is illustrated in Figure 8. The
backbone RMSD between template and model structures is 1.93 Å. Considering the
11
sequence divergence between these ortholog sequences, approximately 25% sequence
identity, this RMSD value is expected and the modeling results, overall, demonstrate that
the MtbSD model obtained is highly satisfactory and can thus be used to infer structure-
activity relationships for this enzyme (Fig. 8).
Discussion
Kinetic Mechanism of MtbSD - The families of intersecting double-reciprocal plots
observed in the initial velocity for the forward direction for both substrates are consistent
with a sequential mechanism and thus both substrates must attach to the enzyme forming a
ternary complex before any product is released. Accordingly, a ping-pong mechanism
could be discarded. In addition, a rapid equilibrium ordered mechanism was ruled out
because the families of double reciprocal plots obtained intersect on the left of the y-axis
(Fig. 2A and Fig. 2B).
The steady-state kinetic constant values given here are larger than the apparent
kinetic parameters we have previously reported [10], whose data were collected in Tris-HCl
100mM, pH 7.0. We verified a significant reduction in the catalytic rate for this enzyme
when the assay mix was buffered using Tris-HCl instead of potassium phosphate (data not
show). In addition, here we report true steady-state kinetic parameters instead of apparent
kinetic parameters.
Determination of steady-state kinetic parameters for Pisum sativum SD [23] in 100
mM potassium phosphate, pH 7.4, yielded values of 340 ± 40
μ
M for K
DHS
, which is
approximately 8-fold larger than the value for MtbSD (44
μ
M), and 4.3 ± 0.5
μ
M for
12
K
NADPH
, which is approximately 8-fold smaller than the value here reported for MtbSD (34
μ
M).
The first step in determining the mechanism of action of an enzyme is to establish
its kinetic mechanism, the order of substrate binding and release of products. This is most
commonly achieved by product inhibition studies in which we observe the patterns
obtained when substrate concentrations are varied at fixed levels of other substrate(s) in the
presence of products, yielding information not available from initial velocity studies.
Product inhibition data for MtbSD (Table 1) were similar to those for P. sativum
SD. For the reaction of P. sativum SD [23], NADPH was found to give linear competitive
inhibition with NADP
+
as variable substrate, and NADPH was a noncompetitive inhibitor
versus SHK at low NADP
+
concentrations. The results for SHK inhibition versus DHS
indicated slightly noncompetitive inhibition, thereby suggesting that the mechanism is
ordered [23]. Product inhibition data presented here suggest two possible kinetic
mechanisms for MtbSD, steady-state ordered and rapid equilibrium random with two dead-
end ternary complexes (MtbSD-NADPH-SHK and MtbSD-NADP
+
-DHS). An application
of primary deuterium isotope effects is to distinguish among possible kinetic mechanisms.
Accordingly, to differentiate between steady-state ordered and rapid equilibrium random,
analysis of primary kinetic isotope effects was carried out.
Measurements of the steady-state primary deuterium kinetic isotope effects on
apparent V/K for one substrate were carried out at various cosubstrate concentrations (Fig.
3). Based on the mechanistic deductions from isotope effects for multireactant enzymes
developed by Cook and Cleland [24], the primary isotope effect on
D
(V/K
app
)
B
is
independent of the concentration of A and
D
(V/K
app
)
A
decreases as the concentration of B
13
increases and reaches a limiting value of 1.0 at infinite concentration of B for a steady-state
ordered mechanism. A and B refer to, respectively, the first and second substrates to bind to
the enzyme. The kinetic isotope effects on V/K results (Fig. 3) are in agreement with a
steady-state ordered bi-bi mechanism with DHS binding first followed by NADPH binding
to MtbSD active site (Fig. 9). In addition, the DHS substrate is sticky, hence it reacts to
give SHK as fast as, or faster than, it dissociates from the enzyme [24]. Based on initial
velocity patterns, product inhibition, analysis of Haldane relationships and isotope-
exchange studies, an ordered kinetic mechanism with NADPH binding first has been
proposed for P. sativum SD [23]. Initial velocity and dead-end inhibition studies showed
that the kinetic mechanism is ordered bi-bi for mouse class II alcohol dehydrogenase with
coenzyme binding first [25]. The NAD-dependent mouse class II alcohol dehydrogenase is
structurally homologous to E. coli SD dinucleotide-binding domain [20].
Stereospecificity and Rate-limiting Steps - The primary deuterium kinetic isotope
effects indicate that the C
4
-proS hydride (B side) is transferred to DHS in the oxy-reduction
reaction catalyzed by MtbSD. In contrast, E. coli SD has been shown to transfer the C
4
-
proR hydride (A-side) of NADPH [26], in agreement with the crystal structure of the
enzyme [20]. Isotope effects on V report on events following formation of the ternary
complex capable of undergoing catalysis, which include the chemical steps, possible
enzyme conformational changes, and product release [27]. Isotope effects on V/K report on
steps in the reaction mechanism from the binding of the isotopically labeled substrate to the
first irreversible step, usually the release of the first product. The apparent classical limit for
primary deuterium kinetic isotope effects on the maximal velocity is around 8 if carbon-
hydrogen bond cleavage is the rate-determining step, even though values as small as 2 have
been used, in a less rigorous practice, as evidence for rate-limiting steps [27]. The primary
14
kinetic isotope effect values of 1.8 for
D
V
DHS
and 1.5 for
D
V
NADPH
using [4S-
2
H]NADPH as
reductant indicate that the hydride transfer is only partly rate limiting for MtbSD enzyme
catalysis. Isotope-insensitive steps such as enzyme isomerization and/or product release are
probably contributing to the rate-limiting steps of the MtbSD reaction. The
D
(V/K
app
)
NADPH
and
D
(V/K
app
)
DHS
values indicate that substrate binding steps make a small contribution to
the rate-limiting steps.
Steady-state solvent isotope effects were evaluated to assess the contribution of
solvent proton transfer to a step in the enzymatic mechanism. The effect on V arises from
solvent-exchangeable protons being transferred during catalysis. The values for solvent
isotope effect on V for DHS (1.5) and NADPH (1.3) indicate that solvent-exchangeable
protons being transferred during catalysis is only partly rate limiting in the overall reaction.
No solvent isotope effect on V/K was observed for DHS and NADPH.
The number of protons transferred during the oxy-reduction reaction catalyzed by
MtbSD was determined by the proton inventory technique. Measurements of V in different
isotopic solvent mixtures (V relative versus mol fraction of D
2
O) showed a linear
relationship (Fig. 4A – inset), suggesting that a single proton is transferred in the step that
exhibits the solvent isotope effect [28]. A similar result was observed for the M.
tuberculosis NADPH-dependent mycothione reductase [29].
Chemical Mechanism of MtbSD - Double isotope effect studies are able to
distinguish whether two different isotopic substitutions affect the same or different
chemical steps. Assuming that the primary deuterium kinetic isotope effect is expressed
only on the hydride transfer reaction (Fig. 5A and 5B – insets) and solvent isotope effects
affect only the alkoxide protonation, double isotope effects can distinguish whether these
15
isotopic substitutions affect the same or different chemical steps (Fig. 5A and 5B). Theory
predicts that if protonation and hydride transfer occur in the same transition state, the
primary isotope effects will be larger or unchanged with D
2
O as compared to H
2
O. On the
other hand, if hydride transfer and protonation occur in distinct steps, the primary isotope
effects will be smaller with D
2
O as solvent, as proton transfer will become more rate
limiting [30, 31]. The increased values for the primary isotope effects on V/K measured in
D
2
O (
D
V/K
DHS(D
2
O)
= 1.3 and
D
V/K
NADPH(D
2
O)
= 1.5) as compared to values measured in H
2
O
(
D
V/K
DHS
= 1.0 and
D
V/K
NADPH
= 1.4) are in agreement with both hydride and proton
transfer taking place in the same step (concerted mechanism), thus implying a single
transition state. It should be pointed out that use of D
2
O allowed the observation of a
significant value for
D
V/K
DHS
(1.3), even though this substrate is sticky and reaction
proceeds through a steady-state ordered kinetic mechanism.
The
β
-elimination catalyzed by bovine liver crotonase is concerted [32], as well as
are the reactions catalyzed by pig liver acyl-CoA dehydrogenase [33] and isocitrate
dehydrogenase [34]. Concerted reactions are a common strategy utilized by enzymes to
avoid unstable intermediates, such as some enolate anions [35, 36].
The role of acid/base chemistry in the mechanism of MtbSD was determined using
the pH dependence of k
cat
and k
cat
/K
m
for DHS and NADPH in pH range of 5.5 pH 10.0
(Fig. 6). In this experiment the k
cat
for both substrates are very similar and decrease at high
pHs with a slope of -1 (Fig. 6A), indicating that deprotonation of a single residue with
apparent pK
a
value of 8.9 ± 0.1 abolishes MtbSD catalytic activity. The k
cat
/K
m
profiles for
DHS and NADPH show that deprotonation of a single ionizable group with pK
a
value of
9.1 ± 0.1 abolishes DHS and NADPH binding (Fig. 6B and 6C). It is likely that the
16
ionization behavior of the same group is being observed in the pH profiles for k
cat
and
k
cat
/K
m
, and the difference in the pK values may reflect perturbation of the pK value upon
substrate(s) binding to MtbSD. The pK values for k
cat
and k
cat
/K
m
lie in the normal pK range
for
ε
-amino of lysine, thiol of cysteines and phenolic hydroxyl of tyrosine amino acids. The
pH-rate profiles of H. influenzae SD for the reverse reaction indicated that a group with pK
value of ~ 8.1 needs to be deprotonated for activity, and site-directed mutagenesis results
were consistent with Lys-67 playing a key role in catalysis [37].
Based on the double isotope effects and pH-rate profiles, we propose a chemical
mechanism for MtbSD (Fig. 10), in which hydride transfer and solvent proton transfer are
concerted, and an amino acid residue with pK
a
value of 8.9 is involved in catalysis.
Analysis of the structure of E. coli SD shows that DHS is positioned for
stereospecific reduction to SHK and the side chain of Lys65 has been proposed to hydrogen
bond to the C-4 hydroxyl group of DHS/SHK (E.coli SD numbering) [20]. This amino acid
residue has been proposed to be the acid/base catalytic group that donates a proton to the
carbonyl of DHS during reduction and that removes a proton during oxidation of SHK [20].
In agreement, previous studies of P. sativum SD showed that substrate-like inhibitors of the
enzyme require a C-4 hydroxyl, whereas either a C-5 hydroxyl or carboxylate group is
needed for strong binding [38]. Analogs of DHS substrate that lack the C-4 and C-5
hydroxyls were used to demonstrate the role of both C-5 hydroxyl and C-4 hydroxyl on the
substrate specificity for the E. coli SD, and the results showed that the C-4 hydroxyl has a
very significant effect on the specificity of the substrate [39]. It has been suggested that the
C-4 hydroxyl group hydrogen bonds to a charged group in the E. coli SD active site based
on an estimation of the binding energy given by k
cat
/K
m
[39]. However, it should be kept in
17
mind that participation of a lysine side chain residue in the chemical reaction catalyzed by
E. coli SD is based on a molecular model, since the crystal structure is for E. coli SD-
NADP
+
binary complex, and E. coli SD transfers the C
4
-proR hydride (A-side) of NADPH.
Based on pH rate profiles of P. sativum SD, it has been postulated that a group with pK
a
value of 9.4, possibly or an
ε
-amino group of lysine, binds the carboxylate ion of the
substrate, while a group of pK
a
8.6, possibly a sulphydryl residue, interacts with the C-4
hydroxyl group of DHS/SHK [40]. The crystal structure of M. jannaschii SD has been
determined and the residues involved in DHS binding and its catalytic reduction were
identified, amongst them Lys70 (M. jannaschii SD numbering) [22]. The side chain of
Asp102 of E. coli SD has also been suggested to form a hydrogen bond to C-4 hydroxyl
group of the substrate [20]. In addition, analysis of the crystal structure of H. influenzae SD
and site-directed mutagenesis studies have shown that Asp103 and Lys67 (H. influenzae
SD numbering) may function as a catalytic pair involved in acid/base catalysis for this
enzyme [37]. However, the pH-rate profiles for MtbSD did not show any residue whose
protonation would abolish enzyme activity at low pH values. However, it could be argued
that the lowest pH value (5.0) of the pH-rate profiles presented here would not allow
detection of carboxylic groups of aspartate side chains since its pK value is approximately
4. Notwithstanding, the pK for the carboxylate group of DHQ/SHK is approximately 4.1
and would be a difficult task to discriminate between substrate and/or amino acid ionizing
groups. We have previously shown by analysis of multiple sequence alignment of MtbSD,
E. coli SD, H. influenzae SD-like, and M. jannaschii SD that Lys69 and Asp105 (M.
tuberculosis numbering) are conserved [10]. The pH-rate profiles described here show
participation of a group whose deprotonation abolishes binding (pK
a
= 9.1) and catalytic
18
activity (pK
a
= 8.9), consistent with participation of Lys69. At any rate, site directed
mutagenesis of these residues and measurements of the steady-state kinetic parameters for
MtbSD enzyme reaction are currently underway to evaluate the role, if any, of Lys69 and
Asp105 may play in binding and/or catalysis.
MtbSD Structure - The MtbSD structure belongs to the homologous superfamily
denominated NAD(P)-binding Rossmann-fold domain [41]. The SDH family provides a
new example of a protein family displaying the dinucleotide binding-fold, without
significant sequence homology with other Rossmann-fold families [20]. The structure has
an elongated shape with two domains (Fig. 8). Each domain has an α/β architecture. The
domains are linked by an α-helix and a turn that keep them together and form a deep groove
into which the NADP cofactor binds. The NADP-binding domain, in the C-terminus,
adopts a nearly canonical Rossmann fold, a six-stranded parallel
β
-sheet with six
α
-helices,
three on each side of the
β
-sheet. The N-terminus consists of a six-stranded
β
-sheet and six
α
-helices, however the arrangement is irregular, since the near central
β
5 strand is in an
antiparallel orientation with respect to the other strands [20].
In the analysis of the E. coli SD structure utilized as template, the substrate-binding
site was identified by the position of the nicotinamide ring of the cofactor [20]. The
following residues conserved in the SD family were identified: Ser14, Ser16, Lys65,
Asn86, Thr101, Asp102, and Gln244 [20]. Both carboxyl oxygens from DHS form two
hydrogen bonds to the protein using probably the conserved serine residues at positions 14
and 16 [20]. These residues were identified in the MtbSD model (Ser18, Ser20, Lys69,
Asn90, Thr104, Asp105, and Gln243; MtbSD numbering), and they are present in
19
equivalent positions as showed in Fig. 11. The conserved residues Ser14, Ser16 and Tyr215
have been suggested to be involved in substrate carboxylate binding for E. coli SD [20].
These residues were found in the MtbSD model (Fig. 11) to be in equivalent positions and
appear to play a role in binding of the carboxylate group of DHS/SKH.
In the multiple sequence alignment between the target (MtbSD) and templates (E.
coli YdiB and E. coli SD) we identified the sequence Gly124-Ser125-Gly126-Gly127-
Thr128-Ala129 as MtbSD sequence pattern (G [A,s,g] G G [A,t] [A,S,g]), corresponding to
the diphosphate-binding loop, which is conserved in the entire SDH family [20] (Fig. 7).
Interaction between the diphosphate-binding loop and the active site in the E. coli SD
occurs through Gly129 and Ala130 residues, which correspond to, respectively, Thr128 and
Ala129 in MtbSD (Fig. 7). For the binding of the nicotinamide in E. coli SD it was verified
that the amide group N-7 of the nicotinamide ring is hydrogen-bonded to the carbonyl
group of Met213 and of the invariant Gly237 [20]. In the MtbSD model, the Gly236 (the
correspondent position for Gly237) is hydrogen-bonded to the amide group N-7. On the
other hand, the MtbSD model presents Ala213 residue in equivalent position of the E. coli
Met213. Notwithstanding, an equivalent bond with the amide group N-7 of nicotinamide
ring is maintained with MtbSD Ala213. The E. coli SD Arg150 and Arg154 play a crucial
role in adenine phosphate binding as they form an “electrostatic clamp” that sandwiches the
phosphate substituent of NADP
+
[20]. In the MtbSD model it was found that Arg149 and
Lys153 residues occupy equivalent positions, and it is thus tempting to suggest that these
residues play a role in adenine phosphate binding in MtbSD.
It has been suggested that NAD
+
binding in E. coli YdiB is favored by the
substitution of residues Thr151 and Arg154 of E. coli SD (E. coli SD numbering) by
Asp158 and Phe160 (E. coli YdiB numbering), respectively [20]. The hydrophobic residue
20
Phe160 creates a neutral environment, which is less discriminating than the basic binding
pocket with Arg154 from the SD structure. The Asn157 and Lys160 residues of NADP
+
-
dependent MtbSD, which correspond to Thr151 and Arg154 of NADP
+
-dependent E. coli
SD [20], occupy equivalent positions and appear to play a similar role in creating a more
discriminating environment in favor of NADP
+
substrate.
The work here presented is, to the best of our knowledge, the first detailed report on
the steady-state velocity patterns, product inhibition, primary deuterium kinetic isotope
effects, solvent kinetic isotope effects, proton inventory, double isotope effects, pH-rate
profiles, and molecular modeling of MtbSD. Initial velocity patterns in the forward
reaction, product inhibition studies, and primary deuterium kinetic isotope effects allowed
us to propose a steady-state ordered bi-bi kinetic mechanism for catalysis, with DHS
binding first followed by NADPH binding to the MtbSD enzyme catalytic site. The primary
deuterium kinetic isotope effects indicated that the C
4
-proS hydride is transferred from the
NADPH co-substrate in a step that is only partly rate limiting. Solvent kinetic isotope
effects demonstrated that proton transfer from the solvent is only partly rate limiting.
Proton inventory results indicated that a single proton is transferred in the solvent-sensitive
step. Double isotope effects showed that transfer of hydride and proton occur in concert.
The pH-rate profiles revealed that a charged group, probably the ε-amino group of Lys69
plays an important role in catalysis and substrate binding. The homology 3D model for
MtbSD identified the probable residues in the substrate binding site, amongst them Lys69,
in agreement with pH studies, and Asp105. These results allowed us to propose a kinetic
and chemical mechanism for MtbSD. The enzyme kinetics and modeling studies provide a
framework on which to base the design of enzyme inhibitors with potential use as
21
antitubercular agents. Site-directed mutagenesis, equilibrium binding spectrofluorimetry,
and pre-steady state experiments are currently underway to improve our understanding of
the mechanism of action of MtbSD.
Acknowledgements
Financial support for this work was provided by Millennium Initiative Program MCT-
CNPq, Ministry of Health-Department of Science and Technology (DECIT)-UNESCO
(Brazil) to D.S.S. and L.A.B. D.S.S. and L.A.B. also acknowledge grants awarded by
CNPq, FINEP, and PRONEX/FAPERGS/CNPq. D.S.S. (CNPq, 304051/1975-06) and
L.A.B. (CNPq, 520182/99-5) are researchers awardees from the National Research Council
of Brazil. We thank Professor John W. Frost, Department of Chemistry of Michigan State
University, for his generous gift of 3-dehydroshikimate substrate.
References
1. C. Dye, S. Scheele, P. Dolin, V. Pathania, M. C. Raviglione, JAMA 282 (1999)
677-686.
2. L. A. Basso, J. S. Blanchard, Adv. Exp. Med. Biol. 456 (1998) 115-144.
3. A. Plabos-Méndez, D.K. Gowda, T.R. Frieden, World Health Organ. 80 (2002)
489-495.
4. CDC (Centers for Disease Control and Prevention), Morb. Mortal. Wkly. Rep. 55
(2006) 301-305.
5. R. J. O’Brien, P. P. Nunn, Am. J. Respir. Crit. Care Med. 162 (2001) 1055-1958.
6. S. T. Cole, R. Brosch, J. Parkhill, T. Garnier, C. Churcher, D. Harris, S. V. Gordon,
K. Eiglmeier, S. Gas, C. E. Barry 3
rd
, F. Tekaia, K. Badcock, D. Basham, D. Brown,
22
T. Chillingworth, R. Connor, R. Davies, K. Devlin, T. Feltwell, S. Gentles, N.
Hamlin, S. Holroyd, T. Hornsby, K. Jagels, A. Krogh, J. McLean, S. Moule, L.
Murphy, K. Oliver, J. Osborne, M. A. Quail, M. A. Rajandream, J. Rogers, S.
Rutter, K. Seeger, J. Skelton, R. Squares, S. Squares, J. E. Sulston, K. Taylor, S.
Whitehead, B. G. Barrell, Nature 393 (1998) 537-544.
7. T. Parish, N. G. Stoker, Microbiology 148 (2002) 3069-3077.
8. J.R. Coggins, C. Abell, L.B. Evans, M. Frederickson, D.A. Robinson, A.W. Roszak,
A.P. Lapthorn, Biochem. Soc. Trans. 31 (2003) 548-552.
9. M. L.Magalhães, C. P. Pereira, L. A. Basso, D. S. Santos, Protein Expr. Purif. 26
(2002) 59-64.
10. I. O. Fonseca, M. L. B Magalhães, J. S. Oliveira, R. G. Silva, M. A.Mendes, M. S.
Palma, D. S. Santos, L. A. Basso, Protein Expr. Purif. 46 (2006) 429-437.
11. G. Ottolina, S. Riva, G. Carrea, B. Danieli, A. F. Buckmann, Biochem. Biophys.
Acta 998, (1989) 173-178.
12. G. A.Orr, J. S. Blanchard, Anal. Biochem. 142, (1984) 232-234.
13. H.M. Berman, J. Westbrook, Z. Feng, , G. Gilliland, T.N. Bhat, H. Weissig,
I.N.Shindyalov, P.E. Bourne, Nucleic Acids Res. 28, (2000) 235-242.
14. S.F. Altschul, T.L. Madden, A.A. Schäffer, J. Zhang, Z. Zhang, W. Miller, D.J.
Lipman, Nucleic Acids Res. 25 (1997) 3389-3402.
15. J.D. Thompson, D.G. Higgins, T.J. Gibson, Nucleic Acids Res. 22 (1994) 4673-
4680
16. A Šali, T.L. Blundell, J. Mol. Biol. 234 (1993) 779-815.
17. M.A. Martí-Renom, A.C. Stuart, A. Fiser, R. Sánchez, F. Melo, A. Šali, Annu. Rev.
Biophys. Biomol. Struct. 29 (2000) 291-325.
23
18. R.A. Laskowski, M.W. MacArthur, D.S. Moss, J.M. Thornton, J. Appl. Cryst. 26
(1993) 283-291.
19. R. Lüthy, J.U. Bowie, D. Eisenberg, Nature 356 (1992) 83-85.
20. G. Michel, A. W. Roszak, V. Sauvé, J. Maclean, A. Matte, J. R. Coggins, M.
Cygler, A. J. Lapthorn, J. Biol. Chem. 278 (2003) 19463-19472.
21. N. Guex, M. C. Peitsch, Electrophoresis 18 (1997) 2714-2723.
22. A.K Padyana, S.K. Burley, Structure 11 (2003) 1005-1013.
23. D. Balinsky, A. W. Dennis, W. W. Cleland, Biochemistry 10 (1971) 1947-1952.
24. P. F. Cook, W. W. Cleland, Biochemistry 20 (1981) 1790-1796.
25. P. Strömberg, S. Svensson, K. B. Berst, B. V. Plapp, J. O. Höög, Biochemistry 43
(2004) 1323-1328.
26. P. Dansette, R. Azerad, Biochimie 56 (1974) 751-755.
27. D. B. Northrop, Biochemistry 14 (1975) 2644-2651.
28. D. M. Quinn, L. D. Sutton, in: P. F. Cook (Ed.), Enzyme Mechanism from Solvent
Isotope Effects, CRC Press, Florida, 1991, pp 73-126.
29. M. P. Patel, J. S. Blanchard, Biochemistry 40 (2001) 5119-5126.
30. J. D. Hermes, C. A. Roeske, M. H. O’Leary, W. W. Cleland, Biochemistry 21
(1984) 5106-5114.
31. J. G. Belasco, J. Albery, J. R. Knowles, J. Am. Chem. Soc. 105 (1983) 2475-2477.
32. Gerlt, J. A. Bioorganic Chemistry: Peptides and Proteins, Oxford University Press,
New York, 1998.
33. B. J. Bahnson, V. E. Anderson, Biochemistry 30 (1991) 5894-5906.
34. B. Pohl, T. Raichle, S. Ghisla, Eur. J. Biochem. 160 (1986) 109-115.
24
35. P. F. Cook, W. W. Cleland, Biochemistry 20 (1981) 1797-1805.
36. A. Thibblin, W. P. Jencks, J. Am. Chem. Soc. 101 (1979) 4963-4973.
37. S. Singh, S. Korolev, O. Koroleva, T. Zarembinski, F. Collart, A. Joachimiak, D.
Christendat, J. Biol. Chem. 280 (2005) 17101-17108.
38. D. Balinsky, D. D. Davies, Biochem. J. 80 (1961) 296-300.
39. T. D. Bugg, C. Abell, J. R. Coggins, Tetrahedron Lett. 29 (1988) 6779-6782.
40. A. W. Dennis, D. Balinsky, Int. J. Biochem. 3 (1972) 93-102.
41. A. G. Murzin, S. E. Brenner, T. Hubbard, C. Chothia. J. Mol. Biol. 247 (1995) 536-
540.
Figure Legends
Fig. 1. Shikimate dehydrogenase-catalyzed reaction.
Fig. 2. Initial velocity patterns for MtbSD with both substrates DHS (A) and NADPH (B) as
variable substrate. Each curve represents varied-fixed levels of the cosubstrate. Both DHS
and NADPH concentrations varied from 5 to 200
μ
M. One unit of enzyme activity (U) is
defined as the amount of enzyme catalyzing the conversion of 1
μ
mol of NADP
+
per
minute at 25 °C.
Fig. 3. Dependence of the apparent
D
V/K
DHS
values on the concentration of the cosubstrate.
The data were fitted to an equation describing a hyperbolic decay, which yielded a limiting
25
value of 1.0 ± 0.03 for the
D
V/K
DHS
. The inset shows that
D
V/K
NADPH
values do not depend
on the concentration of the cosubstrate.
Fig. 4. Solvent isotope effects for MtbSD. (A) DHS as varied substrate, with saturating
concentration of the cosubstrate. Reaction mix contained either 0 () or 90 () atom %
D
2
O. Inset represents the proton inventory on MtbSD, with both substrates at saturating
concentrations. (B) NADPH as varied substrate, with saturating concentration of the
cosubstrate. Reaction mix contained either 0 () or 90 () atom % D
2
O.
Fig. 5. Multiple isotope effects for MtbSD. (A) NADPD as varied substrate, with saturating
concentration of DHS. Reaction mix contained either 0 () or 90 () atom % D
2
O. Inset
represents the primary isotope effect, NADPH () or NADPD () as varied substrate with
saturating concentration of the DHS. (B) DHS as varied substrate, with saturating
concentration of NADPD. Reaction mix contained either 0 () or 90 () atom % D
2
O.
Inset represents the primary isotope effect, DHS as varied substrate with saturating
concentration of NADPH () or NADPD ().
Fig. 6. Dependence of MtbSD kinetics parameters on pH. (A) pH dependence of log k
cat
,
(B) pH dependence of log k
cat
/K
DHS
, and (C) pH dependence of log k
cat
/K
NADPH
.
Experimental data were fitted to equation 7.
Fig. 7. ClustalW multiple sequence alignment between the target (MtbSD) and templates
(E. coli YdiB and E. coli SD). YdiB is more similar to MtbSD than to E. coli SD and
optimal alignment between the three sequences requires the insertion of two gap regions in
26
the C-terminus of MtbSD and E. coli SD sequences (underlined in YdiB).
α
-helices and
β
-
strands are represented as H and S, respectively. This sequence alignment was created
using the following sequences from GeneBank
TM
: M. tuberculosis SD (CAB06186,
residues 1–269), E. coli K12 YdiB (NP 416207, residues 1-288), and E. coli K12 SD
(NP_417740, residues 1-272).
Fig. 8. Stereo ribbon representation of the MtbSD structure. The MtbSD structure is
characterized by a nearly canonical Rossmann fold with a six-stranded parallel
β
sheet (C-
terminal domain on top), and a
α
/
β
N-terminal domain (bottom) with an antiparallel
β
-
strand (
β
5). The α-helices are shown in light gray and the
β
-strands in black.
Fig. 9. Proposed kinetic mechanism for M. tuberculosis shikimate dehydrogenase.
Fig. 10. Proposed chemical mechanism for M. tuberculosis shikimate dehydrogenase-
catalyzed reaction. R represents ribose, adenosine diphosphate, and 2’-phosphate moieties
of NADPH.
Fig. 11. The conserved residues in the nicotinamine ring pocket current in the structures of
both E. coli SD (black) and in the MtbSD (gray). The tyrosine and serine residues involved
in hydrogen-bond with the DHS carboxylate group are in equivalent positions in both E.
coli SD (Tyr215, Ser14, and Ser16; black) and in the MtbSD (Tyr215, Ser18, and Ser20;
gray).
27
Table1: Product Inhibition Patterns for M. tuberculosis Shikimate Dehydrogenase
a
Varied
substrate
Product
inhibition
Inhibition type
b
K
is
(μM)
c
K
ii
(
μ
M)
d
DHS SHK C 62.9 ± 10.5
NADPH SHK NC 256.3 ± 35.4 142.5 ± 74.0
DHS NADP
+
NC 30.4 ± 1.3 28.3 ± 7.6
NADPH NADP
+
C 8.3 ± 1.3
a
At 25°C and 100 mM potassium phosphate, pH 7.3.
b
C = competitive, NC-MT =
noncompetitive mixed-type.
c
K
is
is the slope inhibition constant.
d
K
ii
is the intercept
inhibition constant.
28
Table 2: Kinetic Isotope Effects for M. tuberculosis Shikimate Dehydrogenase
Parameter Isotope Effect*
D
V/K
DHS
1.0 ± 0.03
D
V
DHS
1.8 ± 0.09
D
V/K
NADPH
1.4 ± 0.3
D
V
NADPH
1.5 ± 0.1
D
2
O
V/K
DHS
1.0 ± 0.01
D
2
O
V
DHS
1.5 ± 0.3
D
2
O
V/K
NADPH
1.0 ± 0.03
D
2
O
V
NADPH
1.3 ± 0.2
D
V/K
DHS(D
2
O)
1.3 ± 0.1
D
V
DHS(D
2
O)
2.6 ± 0.01
D
V/K
NADPH(D
2
O)
1.5 ± 0.1
D
V
NADPH(D
2
O)
2.5 ± 0.1
* value ± standard error obtained from fitting the data to
the appropriated equation
29
1
Abbreviations used: DHS, 3-dehydroshikimate; MDR-TB, multidrug-resistant
tuberculosis; MtbSD, Mycobacterium tuberculosis shikimate dehydrogenase; NADP
+
,
oxidized
β
-nicotinamide adenine dinucleotide phosphate; NADPH, reduced
β
-nicotinamide
adenine dinucleotide phosphate; NADPD, deuterated
β
-nicotinamide adenine dinucleotide
phosphate; SD, shikimate dehydrogenase; SHK, D-shikimate; TB, tuberculosis; XDR-TB,
extensively drug-resistant tuberculosis.
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
PARTE III
24
Resultados e Discussão
ARTIGO 1: “Functional shikimate dehydrogenase from
Mycobacterium tuberculosis H37Rv: Purification and
characterization
Obs.: As figuras e tabelas em negrito, mencionadas na discussão a seguir,
estão identificadas de acordo com o artigo a que se referem.
Purificação e Análise da enzima recombinante MtbSD
No artigo intitulado “Functional shikimate dehydrogenase from
Mycobacterium tuberculosis H37Rv: Purification and characterization”, o
protocolo de purificação para a enzima MtbSD utiliza quatro etapas
cromatográficas, onde a proteína recombinante MtbSD foi purificada 8,5 vezes.
Foram necessárias 49g de células E. coli BL21(DE3) com vetor de expressão
pET23a(+) contendo o gene aroE de M. tuberculosis para obtenção de ~ 10 mg
de MtbSD pura. Esse protocolo de purificação e a análise da proteína
recombinante pura foram feitos na dissertação de mestrado da Maria de
Lourdes Borba Magalhães pelo Programa de Pós-Graduação em Biologia
Molecular e Celular da Universidade Federal do Rio Grande do Sul.
Neste protocolo a primeira etapa cromatográfica, utilizando-se da Q-
Sepharose Fast Flow, resultou na maior perda de todo o processo. Apesar
disso, essa etapa foi interessante por possibilitar a redução da quantidade das
proteínas totais em 90% e concentrar a proteína de interesse. Na etapa
subseqüente utilizou-se uma coluna de interação hidrofóbica, a Phenyl
Sepharose High Performance, que resultou em uma taxa de purificação de
quatro vezes e um rendimento de 60% em relação à etapa anterior. A terceira
25
etapa foi feita em gel filtração, através da matriz Sephacryl High Resolution S-
200 e por fim, a coluna de troca aniônica, Mono Q-HR, foi necessária para o
refinamento da purificação da MtbSD (Tabela 1 e Figura1).
A proteína purificada foi ensaiada monitorando a absorbância (340 nm)
da redução do cofator NADP
+
a NADPH por meio da oxidação do D-chiquimato
(SHK) que leva a formação do 3-desidrochiquimato (DHS) (Figura 3). É
possível verificar que a atividade enzimática da proteína purificada foi
linearmente dependente do volume de proteína homogênea à mistura de
reação, mostrando que a velocidade inicial é proporcional à concentração total
de enzima (Figura 2).
A análise do produto da purificação, por espectrometria de massas em
ESI-MS, mostra dois picos (Figura 4). Um pico em 27.207 Da que corresponde
à massa molecular da seqüência primária prevista para a MtbSD, com a perda
da metionina N-terminal; e o segundo pico em 54.150 Da que corresponde à
presença da forma dimérica da MtbSD. O grau de pureza obtido da proteína
pura foi de 98%. A ausência de um pico em 29.413 Da no espectro de massas
confirma que a MtbSD foi separada da enzima EcoliSD (CHAUDURI &
COGGINS, 1985), hospedeiro utilizado na expressão da proteína alvo.
Figura 4: Espectro de massas da MtbSD. O espectro mostra a massa molecular da enzima
na forma monomérica e dimérica. O eixo de valores x representa a massa molecular e o eixo
de valores y a intensidade do pico.
26
A comprovação da natureza dimérica da enzima em solução foi feita por
meio da cromatografia líquida em gel filtração utilizando-se a coluna Superdex-
200 e uma curva de calibração (Figura 5). As SDs de E. coli (MICHEL e cols.,
2003) e H. influenzae (YE e col., 2003) apresentam-se na forma monomérica
em solução, em contrapartida a MjanSD é dimérica (PADYANA & BURLEY,
2003).
FIGURE 5: Determinação da massa molecular da MtbSD em solução. A amostra foi
aplicada na coluna Superdex 200 HR 10/30 e a corrida foi feita em um fluxo de 0.4 mL min
-1
em
FPLC. Os padrões de massa molecular foram Ribonuclease A (13.700 Da), Chymotrypsinogen
(25.000 Da), Ovalbumin (43.000 Da) e Albumin (67.000 Da).
A seqüência dos resíduos de aminoácidos no N-terminal da proteína
purificada, SEGPKKAGVLG, foi determinada pelo método de química
degradativa de Edman, confirmando assim que a proteína pura trata-se da
MtbSD e que realmente ocorre a remoção da metionina N-terminal, um tipo
comum de modificação pós-traducional de proteínas sintetizadas em sistemas
procariotos.
Posteriormente este protocolo de purificação foi otimizado, reduzindo-se
a purificação para três etapas cromatográficas. Em torno de 5 g de células
27
foram necessárias para obtenção de 130,3 mg de proteínas totais, cujo
rendimento de 13,5 % resultou em 1,5 mg de MtbSD pura. Essa preparação
protéica foi utilizada para todos os ensaios de caracterização e de cinética.
Figura 6: Análise por SDS-PAGE das frações oriundas das etapas cromatográficas da
purificação da MtbSD. Canaleta 1: marcador de peso molecular Perfect Marker (GIBCO);
canaleta 2: extrato bruto (40 μg de proteínas); canaleta 3: pool das frações da Phenyl
Sepharose FF (40 μg de proteínas), canaleta 4: pool das frações da Sephacryl-S200 (20 μg de
proteína), canaleta 5: pool das frações da Mono-Q (10 μg de proteína).
Tabela1: Purificação da enzima chiquimato desidrogenase de M. tuberculosis H37Rv.
Etapa de
Purificação
Proteínas
totais (mg)
Unidades
totais (U)
Atividade
específica
(U/mg)
Taxa de
purificação
Rendimento
(%)
Extrato bruto
130,3 156,5 1,2 1 100
Phenyl
Sepharose FF
25,1 113,0 4,5 3,8 72,2
Sephacryl
S200
8,9 81,9 9,2 7,7 52,3
MonoQ
1,5 21,2 14,1 11,8 13,5
Analisando a tabela acima de purificação, a otimização da purificação
ocorreu pela mudança da primeira etapa cromatográfica, onde a coluna de
troca iônica, a Q-Sepharose Fast Flow foi substituida pela coluna de interação
hidrofóbica Phenyl Sepharose Fast Flow, visto que a coluna cromatográfica de
interação hidrofóbica foi bem sucedida no primeiro protocolo de purificação
28
proposto e a coluna Q-Sepharose Fast Flow resultou na maior perda do
processo anterior. Este novo protocolo de purificação reduz uma etapa na
purificação da MtbSD recombinante, melhorando os valores de rendimento e
taxa de purificação.
Os protocolos de purificação estabelecidos aqui, não usam a adição de
seqüências extras para o alcance da proteína purificada; como é o caso da
HinfluSD que utiliza uma cauda de histidinas (YE e cols., 2003). Já foi
demonstrado que esse tipo de artefato pode exercer um efeito negativo em
posteriores estudos da proteína recombinante, alterando à nível de atividade
biológica ou até mesmo alterações estruturais (WOESTENENK e cols., 2004).
Na purificação da enzima MjanSD utilizaram duas etapas cromatográficas e
após mais uma etapa de remoção proteolítica da cauda (PADYANA &
BURLEY, 2003).
Caracterização e Determinação dos Parâmetros Cinéticos Aparentes
O artigo em questão também descreve os parâmetros cinéticos
aparentes no sentido direto e reverso da reação catalisada pela enzima MtbSD
recombinante. É interessante ressaltar que esses resultados para a reação
direta foram os primeiros descritos na literatura para a família das SDs de
procariotos.
Os dados foram coletados e analisados em equação hiperbólica,
indicando que a reação catalisada pela MtbSD segue a cinética de Michaelis-
Menten para todos os substratos (Figura 3). Os valores de K
m
e V
max
encontrados para 3-desidrochiquimato (DHS) foram de 31 μM e 108 U mg
-1
,
respectivamente; e para o cofator NADPH foram de 10 μM e 100 U mg
-1
(Tabela 2). O valor de K
m
para o substrato DHS é bastante inferior ao valor
29
encontrado para a SD purificada de Pisum sativum (PsatSD), cujo valor
descrito é de 340 μM (BALINSKY e cols., 1971). Entretanto, o valor de K
m
encontrado para o cofator NADPH é próximo ao determinado para a enzima
PsatSD em condições semelhantes de pH (4,3 μM). Na reação reversa, os
valores aparentes de K
m
e V
max
para os substratos SHK e NADP
+
foram de
50,18 μM e 18 U mg
-1
; e de 22 μM e 12,9 U mg
-1
, respectivamente (Tabela 2).
O valor de K
m
para o SHK é 10 vezes maior para PsatSD em relação ao de
MtbSD. No entanto, os valores de K
m
para os substratos SHK e NADP
+
das
enzimas MtbSD e EcoliSD são bastante próximos, indicando que as duas
enzimas catalisam a reação reversa com constantes cinéticas semelhantes.
Os valores de V
max
para os substratos DHS e SHK rederam k
cat
e k
cat
/K
m
para esses substratos de 49 s
-1
e 1,6 x 10
6
M
-1
s
-1
, e 8,2 s
-1
e 1,63 x 10
5
M
-1
s
-1
,
respectivamente (Tabela 2).
A constante de equilíbrio aparente (K
eq
) também foi estimada utilizando
os parâmetros cinéticos aparentes em estado estacionário. Por meio da
equação de Haldane o valor de 19,6 foi alcançado para K
eq
. A K
eq
determinada
em pH 7.4 para PsatSD foi de 10,3. É importante ressaltar que a metodologia
utilizada para essa determinação considera o mecanismo cinético como sendo
aleatório em rápido equilíbrio. Nesse caso é assumido que a constante de
dissociação para o segundo substrato não é alterado pela ligação do primeiro
substrato.
O efeito da temperatura na estabilidade da proteína MtbSD foi verificado.
Foram incubadas amostras da proteína recombinante em temperaturas
diferentes (15, 25, 37 e 55°C) e alíquotas foram retiradas em tempos diferentes
de incubação para realização da atividade enzimática (25°C). A MtbSD se
30
mostra uma proteína bastante estável nas temperaturas de 15, 25 e 37°C
durante 1 hora de incubação (Figura 4). Entretanto, a atividade enzimática foi
reduzida gradativamente em 1 hora de incubação a 55°C, chegando a 20% da
atividade inicial no último ponto analisado. A estabilidade da MtbSD também é
constatada a temperatura de -20°C, onde a proteína mantém sua atividade
enzimática por pelo menos 1 ano.
Pela análise do gráfico logk verso 1/T (K
-1
), chegou-se a um valor de
35,2 kJ mol
-1
de energia de ativação (E
a
) para MtbSD (Figura 5). É importante
ressaltar que a E
a
calculada pelo gráfico de Arrhenius é um valor aparente e
que o fator pré-exponencial (A) foi considerado como independente da
temperatura no intervalo de temperatura utilizado nos experimentos. Logo,
esse valor para a E
a
pode ser considerado a quantidade mínima de energia
necessária para iniciar a reação química catalisada pela MtbSD em
concentrações de saturação de SHK e NADP
+
testadas. Outra observação feita
foi que o gráfico se apresenta linear, indicando que a etapa limitante de
velocidade não altera quando a temperatura é mudada. É importante ressaltar
que a veracidade dos dados sobre a E
a
da MtbSD é constatada pela
inexistência de queda repentina do gráfico que indicaria uma possível
desnaturação da enzima, consistente com os dados de estabilidade da
proteína.
Análise da Seqüência Proteica da MtbSD
Através de um alinhamento múltiplo de seqüências, a seqüência de
resíduos de aminoácidos da MtbSD foi comparada com outras três seqüências
de SDs, a de E. coli (MICHEL e col., 2003), a de H. influenzae (YE e col., 2003)
31
e M. jannaschii (PADYANA & BURLEY, 2003), cujas estruturas tridimensionais
já foram determinadas por cristalografia (Figura 6).
A maior identidade encontrada foi de 24% entre as seqüências de
MtbSD e EcoliSD, e entre MtbSD e MjanSD. Apesar desse valor baixo,
considerado o limite para realizar estudos de modelagem molecular, vários
resíduos conservados da família das SDs foram encontrados nas seqüências
analisadas.
O sítio de ligação do substrato na EcoliSD, localizado no domínio N-
terminal, foi identificado pelo posicionamento do anel nicotinamida do cofator
NADPH na estrutura resolvida. Este sítio está em um sulco encontrada na
estrutura da EcoliSD, onde a maior parte dos resíduos altamente conservados
da família das SDs estão presentes, como por exemplo, Ser14, Ser16, Lys65,
Asn86, Thr101, Asp102 e Gln244 (numeração pela seqüência de EcoliSD)
(MICHEL e col., 2003). Esses resíduos estão conservados na seqüência
polipeptídica de MtbSD, correspondendo a Ser18, Ser20, Lys69, Asn90,
Thr104, Asp105 e Gln243 (numeração pela seqüência de MtbSD).
O domínio N-terminal (domínio I) da estrutura da MjanSD é responsável
pela ligação do substrato DHS e os resíduos presentes neste sítio,
aparentemente envolvidos nessa ligação, Lys70, Asn91 e Asp106 (numeração
pela seqüência de MjanSD), são resíduos polares invariáveis em SD. Outra
revelação feita pela estrutura da MjanSD, foi o resíduo Gln254 (numeração
pela seqüência de MjanSD), que também aparece fazendo parte da estrutura
deste sítio e está conservado na família das SDs. Foi proposto que esses
resíduos estão envolvidos com a redução catalítica de DHS a SHK pela
MjanSD (PADYANA & BURLEY, 2003). Pelo alinhamento estes resíduos na
32
seqüência da MtbSD correspondem a Lys69, Asn90, Asp105 e Gln244
(numeração pela seqüência MtbSD).
Pelo estudo estrutural da HinflSD, foi sugerido que os aminoácidos
Ser14, Ser16, Lys65, Asn85, Asp102 e Gln245 (numeração pela seqüência de
HinflSD) estão envolvidos no sítio de ligação do DHS (YE e cols., 2003). Pelo
alinhamento múltiplo esses resíduos correspondem aos resíduos mencionados
anteriormente. Na enzima SD-like de H. influenzae foi proposto que os resíduos
conservados Lys67, Asn88, Asp103 e Gln242 estão envolvidos tanto na
catálise quanto na ligação do substrato (SINGH e cols., 2005). Todos os
resíduos mencionados nos trabalhos da HinfluSD e da SD-like de H. influenzae
estão conservados na seqüência de MtbSD e estão presentes nas demais
seqüências analisadas.
Este primeiro trabalho descreve o primeiro protocolo de purificação para
a enzima MtbSD, que é capaz de obter quantidade suficiente de enzima pura
para fazer experimentos de cinética e de cristalografia. Além disso, apresenta
uma caracterização detalhada informando o estado oligomérico e as
constantes cinéticas aparentes para ambas reações, direta e reversa, entre
outras. As constantes cinéticas da reação direta foram as primeiras descritas
na literatura para as SDs de procariotos. Como perspectiva tem-se a
determinação do mecanismo cinético em estado estacionário e pré-
estacionário, os efeitos isotópicos, a mutagênese sítio direcionada e
“salvamento químico” que permitirão a elucidação dos mecanismos cinético e
químico. A estrutura e o conhecimento do modo de ação da MtbSD serão
utilizados como “conhecimento base” para o desenho de inibidores efetivos da
33
via do chiquimato, pela inibição específica da SD. Esses inibidores serão
potenciais agentes anti-TB.
34
ARTIGO 2: “Shikimate Dehydrogenase from Mycobacterium
tuberculosis H37Rv: Kinetic and Chemical Mechanisms
Obs.: As figuras e tabelas em negrito, mencionadas na discussão a seguir,
estão identificadas de acordo com o artigo a que se referem.
Parâmetros de Velocidade Inicial e Inibição pelos Produtos
O primeiro passo para esclarecer o mecanismo de ação de uma enzima
é estabelecer seu mecanismo cinético, determinando assim a ordem de ligação
dos substratos ao sítio ativo da enzima e posteriormente de liberação dos
produtos. O inicio dessa caracterização foi feita pela determinação das
velocidades iniciais e posteriormente plotando esses valores no gráfico de
Lineweaver-Burk para visualização do perfil das famílias de retas alcançadas.
Para determinar os parâmetros cinéticos em estado estacionário e os
padrões de velocidade inicial, a atividade da MtbSD foi determinada variando a
concentração de um dos substratos em algumas concentrações fixas do outro
substrato e vice-versa.
Esses dados foram plotados no gráfico de Lineweaver-Burk, e as
famílias de retas observadas na reação direta para os dois substratos são
consistentes com o mecanismo seqüencial e como conseqüência ocorrendo a
formação de um complexo ternário no decorrer da reação. Com isso foi
descartada a possibilidade do mecanismo ping-pong para a MtbSD, que
apresenta no plote de Lineweaver-Burk uma família de retas paralelas. Nos
ensaios onde DHS foi variado em sete concentrações fixas de NADPH (10, 25,
50, 100 e 200
μ
M), as retas se interceptaram sobre o eixo x e a esquerda do
eixo y (Figura 2A). Quando o substrato NADPH foi variado, seguindo as
35
mesmas condições de ensaio para DHS, as retas se interceptaram abaixo do
eixo x e a esquerda do eixo y (Figura 2B). Com estes gráficos foi descartado
também o mecanismo ordenado em equilíbrio rápido, pois nesse mecanismo
se espera uma família de retas que se interceptam no eixo y.
Os dados coletados foram ajustados na equação geral para mecanismo
seqüencial v = VAB / (K
a
B + K
b
A + K
ia
K
b
+ AB), onde v é a velocidade medida
da reação, V é a velocidade máxima, A e B são as concentrações dos
substratos, K
a
e K
b
são as constantes de Michaelis-Menten dos substratos A e
B e K
ia
é a constante de dissociação para o substrato A. Os valores para as
constantes de Michaelis-Menten dos substratos DHS e NADPH é de 44 ± 2,8
μ
M e 34 ± 2,2
μ
M, respectivamente. A constante catalítica (k
cat
) foi de 77,8 ±
2,3 s
-1
,
alcançando então as constantes de especificidade de 1,8 x 10
6
M
-1
s
-1
(k
cat
/K
DHS
) e de 2,3 x 10
6
M
-1
s
-1
(k
cat
/K
NADPH
). Esses valores são maiores que os
parâmetros cinéticos aparentes apresentados no artigo “Functional shikimate
dehydrogenase from Mycobacterium tuberculosis H37Rv: Purification and
characterization”, onde os dados foram coletados em tampão Tris HCl 100 mM
pH 9,0. Foi verificado um aumento significativo na velocidade catalítica da
MtbSD quando no ensaio de reação utilizou-se tampão fosfato de potássio 100
mM pH 7,3 ao invés de tampão Tris HCl 100 mM pH 9,0. Os parâmetros
cinéticos em estado estacionário para a PsatSD, em tampão fosfato de
potássio 100 mM pH 7,3, são: K
DHS
340 ± 40
μ
M e K
NADPH
4.3 ± 0.5
μ
M
(BALINSKY e col., 1971). Observando esses valores, se constata que K
DHS
para PsatSD é 8 vezes maior e que K
NADPH
é 8 vezes menor em relação aos
valores determinados para MtbSD.
36
O estudo de inibição pelos produtos foi o segundo passo para a
determinação do mecanismo cinético da MtbSD. Os dados coletados foram
ajustados nas seguintes equações: v = VA / [K
a
(1 + I/K
is
) + A] ou v = VA / [K
a
(1
+ I/K
is
) + A (1 + I/K
ii
)], onde v é a velocidade medida da reação, V é a
velocidade máxima, A é a concentração do substrato, K
a
é a constante de
Michaelis-Menten do substrato A, K
is
é a constante de inibição angular, K
ii
é a
constante de inibição linear e I é a concentração do inibidor.
O substrato NADP
+
mostrou-se um inibidor competitivo para NADPH e
não competitivo para DHS. O produto SHK é um inibidor competitivo e não
competitivo para os substratos DHS e NADPH, respectivamente (Tabela 1).
Esse perfil de inibição pelos produtos é compatível com os mecanismos
cinéticos ordenado em estado-estacionário e aleatório em rápido equilíbrio.
Dados similares para a inibição pelos produtos foram descritos para a PsatSD
(BALINSKY e col., 1971).
Esses dados de inibição pelos produtos indicam que no decorrer da
reação dois possíveis complexos não produtivos podem ser formados, o
complexo MtbSD-NADPH-SHK e o complexo MtbSD-NADP
+
-DHS, caso o
mecanismo cinético tratar-se de aleatório em rápido equilíbrio.
Efeitos Cinéticos Isotópicos Primário do Deutério
A diferenciação entre os dois possíveis mecanismos para a MtbSD foi
feita através de experimentos de efeitos cinéticos isotópicos primário. Para isso
foram determinados os parâmetros cinéticos em velocidade inicial variando um
dos substratos em cinco concentrações fixas do outro substrato, utilizando
NADPH ou
β
-nicotinamida adenina dinucleotídeo fosfato deuterado ([4S-
37
2
H]NADPH /NADPD). Os efeitos isotópicos cinéticos primário foram observados
sobre o V/K e o V em relação à reação não marcada.
O efeito isotópico sobre V/K descreve o efeito a partir da ligação do
substrato marcado até a primeira etapa irreversível da reação, que geralmente
está representada pela liberação do primeiro produto. O efeito isotópico sobre
V exibe o efeito nos eventos seguintes após a formação do complexo ternário
produtivo, incluindo as etapas químicas, possíveis isomerizações da enzima
até a liberação dos produtos (NORTHROP, 1975).
Os valores coletados foram ajustados nas seguintes equações: v = VA /
[K (1 + FiE
V/K
) + A (1 + FiE
V
) ou v = VA / [K + A (1 + FiE
V
)], as quais descrevem
o efeito sobre o V e V/K, e apenas sobre V, respectivamente. Os termos Fi
representa a fração isotópica marcada, Evk o efeito isotópico sobre V/K menos
1 e Ev o efeito isotópico sobre V menos 1 (Tabela 2).
O valor do efeito isotópico cinético primário de 1,8 para
D
V
DHS
e de 1,5
para
D
V
NADPH
utilizando [4S-
2
H]NADPH como redutor, indica que a etapa de
transferência do hidreto, que compreende parte da etapa química é
parcialmente limitante para a velocidade catalítica da MtbSD. Isso sugere que
outras etapas insensíveis ao efeito isotópico, tais como isomerização da
enzima e/ou liberação dos produtos, estão envolvidas com as etapas limitantes
de velocidade da reação catalisada pela MtbSD. Os valores baixos para
D
V/K
NADPH
e
D
V/K
DHS
indicam que a etapa de ligação dos substratos até a
formação do complexo irreversível tem uma pequena contribuição na
velocidade de catalise pela MtbSD.
A magnitude dos efeitos isotópicos cinéticos utilizando o [4S-
2
H]NADPH
indica que o hidrogênio proS do C
4
(lado B) do cofator é transferido na forma
38
de hidreto para o substrato DHS na reação de catálise pela MtbSD. Estudo
similar feito com a EcoliSD demonstrou que o hidreto transferido na mesma
reação de oxiredução é o proR do C
4
(lado A) do NADPH (DANSETTE &
AZERAD, 1974), dado que confere com o estudo da estrutura 3D da proteína
EcoliSD (MICHEL e col., 2003).
A dedução de mecanismo cinético utilizando os efeitos isotópicos para
enzimas com multireagentes foi desenvolvida por Cook & Cleland (1981),
utilizando os valores dos efeitos isotópicos cinéticos aparentes sobre V/K para
um substrato em várias concentrações do cosubstrato.
Os valores do
D
(V/K
app
)
NADPH
(1,4 ± 0,3) são independentes da
concentração de DHS, entretanto os valores de
D
(V/K
app
)
DHS
diminuem até o
valor de 1,0 ± 0,03 quando a concentração de NADPH ou NADPD é
aumentada (Figura 3). Segundo as deduções mecanísticas de Cook & Cleland
quando o efeito isotópico primário
D
(V/K
app
)
B
é independente da concentração
de A e o efeito isotópico primário
D
(V/K
app
)
A
diminui com o aumento da
concentração de B, chegando a 1,0 em concentração infinita de B, o
mecanismo cinético da reação em estudo é ordenado em estado estacionário.
A e B refere-se ao primeiro e ao segundo substrato, respectivamente, que se
liga a enzima. Frente a esses dados chega-se a conclusão de que o
mecanismo cinético para a MtbSD é o mecanismo ordenado em estado
estacionário, e que a ordem de ligação ao sítio catalítico da enzima é DHS
primeiro, seguida da ligação do cofator NADPH (Figura 10).
Utilizando os dados dos parâmetros de velocidade inicial, de inibição
pelos produtos, da análise da relação de Haldane e dos estudos isotope-
exchange, o mecanismo cinético ordenado com NADPH ligando primeiro foi
39
proposto para a PsatSD (BALINSKY e col., 1971). A enzima álcool
desidrogenase NAD-dependente (classe II) de camundongo, estruturalmente
homóloga ao domínio de ligação do cofator da EcoliSD (MICHEL e col., 2003),
apresenta o mecanismo seqüencial ordenado bi-bi com o cofator NADPH
ligando antes do outro substrato (Strömberg e col., 2004).
Efeito Isotópico Cinético do Solvente e “Proton Inventory”.
Para analisar a contribuição da etapa de transferência de próton pelo
solvente na velocidade da reação enzimática e para determinar o número de
prótons que são transferidos pelo solvente foram feitos experimentos de efeito
isotópico cinético do solvente e “próton inventory”, respectivamente.
O efeito isotópico cinético do solvente foi determinado medindo as
velocidades iniciais usando concentrações saturantes de um substrato variando
a concentração do outro em H
2
O e em 90% de D
2
O (Figura 4A e 4B). Os
valores coletados foram ajustados nas equações utilizadas para determinação
dos efeitos isotópicos primário. Os efeitos cinéticos isotópicos do solvente
observados sobre o V/K e o V estão mostrados na Tabela 2.
Os valores dos efeitos isotópicos do solvente sobre V (
D
2
O
V) foram
pequenos, atingindo para DHS a magnitude de 1,5 ± 0,3 e para NADPH de 1,3
± 0,2. O efeito sobre V/K (
D
2
O
V/K) não foi observado. Esses dados sugerem
que a transferência de próton pelo solvente não se trata da etapa limitante de
velocidade, e sim um contribuinte parcial para a velocidade da reação
catalisada pela MtbSD.
“Proton inventory” foi determinado utilizando a relação entre a velocidade
inicial em concentrações saturantes de NADPH e DHS e várias concentrações
de D
2
O (20, 40, 60, 80 e 90%). Esta relação foi linear, sugerindo que ocorra a
40
transferência de um único próton pelo solvente na reação de oxiredução
catalisada pela MtbSD (QUINN & SUTTON, 1991). Um resultado similar foi
observado para a enzima micotiona redutase dependente de NADPH de M.
tuberculosis (PATEL & BLANCHARD, 2001) (Figura inserida 4A)
Efeitos Isotópicos Cinéticos Múltiplos
A fim de caracterizar o mecanismo químico da reação catalisada pela
MtbSD, foram feitos os efeitos isotópicos cinéticos múltiplos através da
determinação dos efeitos isotópicos cinéticos do solvente utilizando NADPD
como substrato variado. Deste modo distingui-se entre o mecanismo químico
concertado, onde a transferência do próton e do hidreto ocorrem na mesma
etapa e o mecanismo químico stepwise, onde a transferência do próton e do
hidreto ocorrem em etapas distintas.
A teoria diz que, se caso a protonação e a transferência de hidreto
ocorrerem na mesma etapa de transição, o efeito isotópico primário será maior
ou igual com o uso de D
2
O como solvente quando comparado com a H
2
O. Por
outro lado, se caso a transferência de hidreto e a protonação ocorrerem em
etapas distintas, o efeito isotópico primário será menor com o uso de D
2
O como
solvente, visto que a transferência de próton se tornará limitante na velocidade
da reação (HERMES e col., 1984 e BELASCO e col., 1983).
Os resultados dos efeitos isotópicos cinéticos múltiplos estão mostrados
na tabela 2. O aumento nos valores de todos os efeitos isotópicos primários
medidos em D
2
O indicam o mecanismo químico concertado para a MtbSD
(Figura 5A e 5B). Um dado importante é o efeito isotópico sobre
D
V/K
DHS
, que
não era observado pelo efeito isotópico cinético primário e passou para 1,3 ±
0,1 quando foi determinado em D
2
O como solvente.
41
O mecanismo concertado é uma estratégia comum utilizada pelas
enzimas para evitar a formação de intermediários instáveis (COOK &
CLELAND, 1981 e THIBBLIN & JENCKS, 1979). A
β
-eliminação catalisada pela
crotonase é do tipo concertado (GERLT, 1998), bem como as reações
catalisadas pela acil-CoA desidrogenase (BAHNSON & ANDERSON, 1991) e a
isocitrato desidrogenase (POHL e col., 1986) de fígado suíno.
Estudos de pH
O papel da química ácido/base no mecanismo da MtbSD foi determinado
através das constantes k
cat
e k
cat
/K
m
em diferentes pHs (5,5 – 10,0).
Previamente foi verificada a estabilidade da enzima MtbSD em cada um dos
pHs testados, para a certificação da veracidade dos dados coletados. Em cada
pH testado, as velocidades iniciais foram medidas variando a concentração de
um dos substratos mantendo o outro em concentração saturante. Os dados
coletados foram ajustados na equação log y = log [C/(1 + K
b
/H)] onde y é o
parâmetro cinético aparente, C é valor de y no platô independente do pH, H é a
concentração de H
+
e K
b
é a constante de dissociação aparente para os grupos
ionizados.
Os efeitos do pH na k
cat
para os substratos DHS e NADPH são similares,
diminuindo os valores em pHs elevados com uma inclinação de -1 (Figura 6A).
Isso demonstra que a deprotonação de um único resíduo com pK
a
aparente de
8,9 ± 0,1 anula a atividade catalítica da MtbSD.
A constante k
cat
/K
m
para ambos os substratos sofreu um decréscimo em
altos pHs com uma inclinação de -1, sugerindo que um único grupo
deprotonado com valor de pK
a
aparente
igual a 9,1 ± 0,1, reduz a ligação entre
substrato-enzima (Figura 6B e 6C).
42
Provavelmente é a ionização do mesmo grupo que esta sendo
observado pelo efeito do aumento de pH sobre as constantes analisadas. O
valor de pK
a
para as constantes k
cat
e k
cat
/K
m
encontra-se na faixa normal de
pK
a
para os grupos
ε
-amino da lisina, tiol da cisteína e hidroxila fenólica da
tirosina. Frente aos resultados de efeitos isotópicos cinéticos múltiplos e
estudos de pH foi proposto um mecanismo químico para a MtbSD mostrado na
figura 11.
A análise da estrutura da EcoliSD propõe que a cadeia lateral do resíduo
Lys65 estabelece uma ponte de hidrogênio ao grupo hidroxila do C-4 do
DHS/SHK (MICHEL e col., 2003). Os autores propuseram ainda que tal resíduo
seja o grupo catalítico ácido/base que doa um próton para o grupo carbonil do
substrato DHS durante a sua redução (MICHEL e col., 2003). Compostos
análogos ao substrato DHS que não continham as hidroxilas no C-4 e C-5
foram utilizadas para demonstrar o papel dessas hidroxilas na especificidade
da enzima EcoliSD pelo substrato. Os resultados mostraram que a hidroxila do
C-4 tem um efeito bastante significativo na especificidade da enzima pelo
substrato (BUGG e col., 1983). A avaliação feita na energia de ligação através
do k
cat
/K
m
sugere que a hidroxila no C-4 estabelece uma ponte de hidrogênio
com um grupo carregado no sítio ativo da EcoliSD (BUGG e col., 1983). É
interessante ressaltar que a participação da cadeia lateral da lisina na reação
química catalisada pela EcoliSD está baseada em um modelo molecular, e a
estrutura determinada para EcoliSD por cristalografia é o complexo binário
enzima-NADP
+
e que a transferência do hidreto nessa reação é dada pelo
hidrogênio proR do C-4 do cofator. A cadeia lateral do resíduo Asp102 da
43
EcoliSD se propõe que também esteja envolvida com a hidroxila no C-4 do
DHS por meio de uma ponte de hidrogênio (MICHEL e col., 2003).
Estudos prévios da PsatSD mostraram que inibidores do tipo substrato-
like necessitam da hidroxila no C-4, ao passo que tanto uma hidroxila quanto
um grupo carboxila no C-5 é necessário para uma ligação forte entre inibidor-
enzima (BALINSKY & DAVIES, 1961). Os efeitos de pH sobre a velocidade da
PsatSD sugerem que uma cisteína ou um grupo
ε
-amino está protonado para
interagir com o grupo hidroxila no C-4 do DHS/SHK (DENNIS & BALINSKY,
1972).
A estrutura da MjanSD foi determinada e os resíduos envolvidos na
ligação e na sua redução catalítica do DHS foram identificados, dentre eles
está a Lys 70 (PADYANA & BURLEY, 2003). A HinflSD tem sua estrutura
resolvida pela análise cristalográfica e estudos de mutagênese sítio direcionada
mostrou que o par Asp103 - Lys67 pode estar envolvido no mecanismo de
catálise desta enzima (SINGH e col., 2005).
Em nossos estudos de efeitos de pH na reação catalisada pela MtbSD
não foi identificado qualquer resíduo cuja protonação anule a atividade
enzimática em pH baixo. Entretanto, os efeitos dos pHs abaixo de 5,5 não
foram testados e com isso não foi permitida a detecção do grupo carboxila
presente na cadeia lateral do aspartato, visto que o valor de pK
a
para esse
grupo é de aproximadamente 4. Os testes em pHs inferiores a 5,5 foram
evitados pelo fato de que o pK do grupo carboxila do DHS/SHK é ~ 4,1 e isso
dificultaria a interpretação dos resultados.
Anteriormente foi feita uma análise da seqüência de MtbSD através do
alinhamento múltiplo de seqüências utilizando: a EcoliSD, a HinflSD-like e a
44
MjanSD, cujas estruturas já foram determinadas (FONSECA e col., 2006).
Dentre os resíduos identificados como conservados para a família das SDs
foram identificados os resíduos Lys69 e Asp105 da MtbSD que correspondem
aos resíduos discutidos nas demais estruturas (FONSECA e col., 2006). Os
dados descritos aqui no estudo de pH para MtbSD mostrou a participação de
um grupo cuja deprotonação desfavorece a ligação da MtbSD ao substrato
(pK
a
= 9.1) e reduz a atividade catalítica (pK
a
= 8.9), que é consistente com a
participação da Lys69 identificada na seqüência primária da enzima alvo.
Modelagem por Homologia
O primeiro passo para execução da técnica de modelagem por
homologia é buscar um molde para os estudos. Foram encontrados três
candidatos: a EcoliSD de (PDB ID: 1NYT) (MICHEL e col., 2003), a
quinato/chiquimato desidrogenase (QSD) de E. coli (PDB ID: 109B) (MICHEL e
col., 2003) and a MjanSD (PDB ID: 1NVT) (PADYANA & BURLEY, 2003). O
resultado do “BLASTp” identificou a MjanSD (PDB ID: 1NVT) como melhor
modelo para a MtbSD. Entretanto, as coordenadas da estrutura não estavam
disponíveis para o estudo. O segundo candidato pelos dados obtidos no
Blastp” foi a QSD de E. coli (PDB ID: 109B), entretanto sua função é diferente
da exercida pela MtbSD e além disso a QSD tem a capacidade de utilizar como
cofator tanto o NADH quanto o NADPH. Outro ponto desfavorável em relação
ao molde QSD é a inserção de dois gaps na seqüência da MtbSD observados
no alinhamento múltiplo de seqüências (Figura 7). Esses segmentos que não
apresentam equivalência no molde são regiões difíceis de resolver na estrutura
a ser modelada.
45
Frente a essas justificativas, a EcoliSD (PDB ID: 1NYT) foi eleita como
molde para a MtbSD, pois apresentam a mesma função biológica, a sua
estrutura foi resolvida em 1,5Å (MICHEL e col., 2003) e a identidade entre as
duas seqüências é de 25%, limite permitido para os estudos de modelagem
estrutural de proteínas sejam feitos (ALTSCHUL e col., 1997).
Partindo desse molde, dez modelos foram construídos para a MtbSD.
Todos os modelos foram avaliados pelos programas PROCHECK
(LASKOWSKI e col. 1993) (Tabela 4) e VERIFY 3D (LUTHY e col., 1992) para
eleger o melhor modelo. Dos 269 resíduos de aminoácidos da MtbSD na
melhor estrutura modelada, 90,5% dos resíduos estão localizados na região
mais favorável no gráfico de Ramachandran (Figura 8 e Tabela 3). O RMSD
da cadeia principal entre as estruturas do molde e da modelada é de 1,93 Å.
Considerando as divergências entre as duas seqüências utilizadas no estudo, o
valor para RMSD é plausível e os resultados gerais da estrutura modelada
mostram que o molde eleito é bastante satisfatório e pode ser utilizado para
analisar a relação estrutural e funcional da MtbSD.
A estrutura da MtbSD pertence a superfamília denominada domínio
Rossmann-fold“ de ligação a NAD(P) (MURZIN e col., 1995). A estrutura
apresenta uma forma alongada com dois domínios, que apresentam uma
arquitetura α/β (Figura 9). Esses domínios estão ligados por uma α-hélice e
uma volta que os mantém juntos e formando um sulco onde o cofator NADPH
liga-se. O domínio de ligação do cofator NADPH, localizado na porção C-
terminal, apresenta uma estrutura do tipo “Rossmann-fold“, que compreende
em seis fitas
β
paralelas e seis
α
hélices, onde cada três
α
hélices estão
dispostas sobre cada lado do leque central de fitas
β
. A porção N-terminal é
46
bastante semelhante ao domínio C-terminal descrito, entretanto ocorre um
arranjo irregular que compreende a orientação antiparalela da fita
β
-5 em
relação às demais fitas (MICHEL e col., 2003).
Na análise da estrutura da EcoliSD, utilizada como molde, foi identificado
o sítio de ligação dos substratos pela posição do anel nicotinamida do cofator,
e neste sulco os resíduos de aminoácidos, Ser14, Ser16, Lys65, Asn86,
Thr101, Asp102 e Gln244 foram identificados. Ambos os oxigênios carboxil do
substrato DHS formam pontes de hidrogênio com o sítio ativo da proteína,
provavelmente pelos resíduos conservados de serina nas posições 14 e 16
(MICHEL e col., 2003). Tais resíduos foram identificados no modelo para
MtbSD, Ser18, Ser20, Lys70, Asn91, Thr106, Asp107 e Gln263, em posições
estruturalmente equivalentes (Figura 12). Na estrutura da EcoliSD foi
identificado o resíduo Tyr215 que também está envolvido na ligação do
substrato DHS no sítio ativo da enzima através de ponte de hidrogênio. No
modelo da MtbSD o resíduo equivalente é Tyr215 (Figura 12).
No alinhamento múltiplo de seqüências entre o alvo (MtbSD) e os
moldes (EcoliQSD and EcoliSD) foi identificada a seqüência Gly124-Ser125-
Gly126-Gly127-Thr128-Ala129 como a seqüência padrão da MtbSD (G [A,s,g]
G G [A,t] [A,S,g]), correspondendo ao loop de ligação ao difosfato, que está
conservado na família das SDs (MICHEL e col., 2003) (Figura 7).
Para a ligação da nicotinamida na EcoliSD foi verificado que o N-7 do
grupo amida do anel nicotinamida é ligado por pontes de hidrogênio ao grupo
carbonil da Met213 e da Gly237 (MICHEL e col., 2003). A Gly236 (posição
correspondente para a Gly237) é conservada no modelo da MtbSD,
estabelecendo o mesmo tipo de interação com o cofator; em contrapartida na
47
posição equivalente a Met213 encontra-se a Ala213 que mantém a mesma
natureza de ligação com o N-7 do anel (Figura 7).
As argininas 150 e 154 na EcoliSD têm um papel crucial na ligação do
sítio ativo ao fosfato da adenine pela formação de um “grampo eletrostático”
(electrostatic clamp) que envolve o fosfato (MICHEL e col., 2003). No modelo
da MtbSD são encontrados os resíduos correspondentes Arg149 e Lys153.
Provavelmente no modelo em estudo estes resíduos têm o mesmo papel, visto
que aparecem pontes de hidrogênio entre eles e o fosfato da adenina (Figura
7).
O resíduo de Ser189 na EcoliSD é ligado ao oxigênio do anel da ribose
fosfato do cofator por ponte de hidrogênio. Analisando o modelo MtbSD essa
interação é dada pelo resíduo de Val189. A interação entre o grupo fosfato da
ribose e o sítio ativo está sendo feito por ponte de hidrogênio com Asn149 e
Asn150 na Ecoli SD e na MtbSD, respectivamente (Figura 7).
A interação entre a porção difosfato do cofator e o sítio ativo da EcoliSD
ocorre através da Gly129 e da Ala130, que corresponde a Thr128 e a Ala129
no modelo em estudo, respectivamente.
Michel e col. (2003) mostraram que a ligação do NAD
+
na EcoliQSD é
favorecida pela substituição dos resíduos Thr151 e Arg154, presentes na
EcoliSD, pelos resíduos Asp158 e Phe160, respectivamente. A explicação para
tal, é o ambiente hidrofóbico criado pelo resíduo apolar Phe160, que reduz a
especificidade pelo NADPH e permitindo assim, que tanto NADH quanto
NADPH sejam utilizados como cofatores pela EcoliQSD. Em posição
equivalente aos resíduos Thr151 e Arg154 na estrutura da EcoliSD estão os
resíduos Asn157 e Lys160 na estrutura modelada da MtbSD. Tanto no molde
48
quanto na estrutura modelada, esses resíduos estabelecem ponte de
hidrogênio e é essa interação que resulta em uma orientação similar das
cadeias laterais e como conseqüência um ambiente equivalente para a ligação
do NADPH (Figura 13).
O trabalho apresentado é a primeira descrição detalhada sobre os
mecanismos cinético e químico para a MtbSD, propondo um resíduo envolvido
na catálise e na etapa de ligação dos substratos.
Como perspectivas próximas há os estudos de mutagênese sítio
direcionada para confirmar a importância do resíduo Lys69 na ligação dos
substratos e na velocidade de catálise pela MtbSD e os estudos de
fluorescência de proteína que fornecerão as constantes de associação e
dissociação em equilíbrio para ambos substratos. A etapa subseqüente será a
cinética em estado pré-estacionário que possibilitará a determinação de
constantes de velocidade de primeira e segunda ordens o que possibilitará um
melhor detalhando do mecanismo cinético.
Estes estudos juntamente com os estudos estruturais do sítio ativo da
enzima fornecerão a base sólida para o desenho racional de inibidores para a
MtbSD, que poderão representar possíveis agentes quimioterápicos para
tratamento da tuberculose.
49
Conclusões
A proteína recombinante purificada é a MtbSD na forma dimérica;
A reação catalisada é reversível, com uma Keq de 19,6, uma Ea = 35,2
kJ mol-1 e segue a cinética de Michaelis-Menten;
O mecanismo cinético é ordenado em estado estacionário, com DHS
ligando primeiro à enzima;
As transferências de hidreto e próton contribuem modestamente para
determinação da velocidade da reação e ocorrem de forma concertada;
Há catálise ácido/base na reação e a protonação de um grupo é
essencial para a catálise e a ligação dos substratos;
Todos os resíduos sugeridos importantes para catálise na SD de E. coli
estão conservados e em posições semelhantes no sítio ativo da MtbSD.
50
Referências
[1]. ALTSCHUL, S.F., MADDEN, T.L., SCHÄFFER, A.A., ZHANG, J.,
ZHANG, Z., MILLER, W., LIPMAN, D.J. (1997) Nucleic Acids Res. 25,
3389-3402
[2]. BALINSKY, D., DAVIES, D. D. (1961) Biochem. J. 80, 296-300
[3]. BALINSKY, D., DENNIS, A.W., CLELAND, W.W. (1971) Biochemistry
10 1947–1952.
[4]. BASSO, L.A., SILVA, L.H.P., FETT-NETO, A.G., AZEVEDO, W.F.,
MOREIRA, I.S., PALMA, M.S., CALIXTO, J.B., ASTOLFI, S., SANTOS,
R.R., SOARES, M.B.P., SANTOS, D.S. (2005) Mem Inst Oswaldo Cruz
100, 475-506.
[5]. BELASCO, J. G., ALBERY, J., AND KNOWLES, J. R. (1983) J. Am.
Chem. Soc. 105, 2475-2477.
[6]. BUGG, T. D., ABELL, C., COGGINS, J. R. (1988) Tetrahedron Lett. 29,
6779-6782
[7]. BENACH, J., LEE, I., EDSTRON, W., KUZIN, A. P., CHIANG, P.,
ACTON, T. B., MONTELIONE, G. T. AND HUNT J. F. (2003) The
Journal of Biological Chemistry 278, 21, 19176-19182.
[8]. BENTLEY, R. (1990) Critical Reviews in Biochemistry and Molecular
Biology 25, 307-384.
[9]. BLOOM, B.R. & MURRAY, C.J.L. (1992) Science 257, 1055-1064.
[10]. BRENNAN, P.J. (1997) FEMS Immunology and Medical Microbiology
18, 263-269.
[11]. CHAUDHURI, S. AND COGGINS J. R. (1985) Biochemical Journal
226, 217-223.
[12]. COLE, S.T., BROSCH, R., PARKHILL, J. et al. (1998) Nature 393,
537–544.
[13]. COOK, P. F., CLELAND, W. W. (1981) Biochemistry 20, 1790-1796
[14]. DANSETTE, P., AZERAD, R. (1974) Biochimie 56, 751-755
[15]. DANIEL, T.M. Captain of death: the story of tuberculosis. 1
th
ed. New
York. University of Rochester Press, 1997.
[16]. DENNIS, A. W., BALINSKY, D. (1972) Int. J. Biochem. 9, 93-102
51
[17]. DORMANDY, T. (2002) The white death: a history of tuberculosis, New
York University Press.
[18]. DUNCAN, K. (2003) Tuberculosis 83, 201-207.
[19]. DUNLAP, N. E., BASS, J., FUJIWARA, P., HOPEWELL, P.,
HORSBURGH JR., C. R., SLAFINGER, M. AND SIMONE, P. M. (2000)
American Journal of Respiratory and Critical Care Medicine 161, 1376-
1395.
[20]. ENARSON, D.A. & MURRAY, J.F. (1996) Global epidemiology of
tuberculosis. In: Tuberculosis (Rom, W.M. & Garay, S., Eds.), pp. 57-75.
Little, Brown and Co., Boston, MA.
[21]. ENSERINK, M. (2001) Science 293, 5528, 234-236.
[22]. ESPINAL M. A. (2003) Tuberculosis 83, 44-51.
[23]. FÄTKENHEUER, G., TAELMAN, H., LEPAGE, P., SCHWENK, A. &
WENZEL, R. (1999) Diagn. Microbiol. Infect. Dis. 34, 139-146.
[24]. FONSECA, I. O., MAGALHÃES, M. L. B., OLIVEIRA, J. S., SILVA, R.
G., MENDES, M. A., PALMA, M. S., SANTOS, D. S., BASSO, L. A.
(2006) Protein Expr. Purif. 46, 429-437
[25]. GLICKMAN, M.S. & JACOBS, W.R. Jr. (2001) Cell 104, 477-485.
[26]. HERMES, J. D., ROESKE, C. A., O’LEARY, M. H., AND CLELAND, W.
W. (1984) Biochemistry 21, 5106-5114.
[27]. KNIGGE, A., MORR, H. AND KILLIAN, A. (2000) Tuberculosis and
HIV/AIDS. Epidemics: Opportunities for cross support. Disease Control
(Tuberculosis) and Health Promotion AIDS Control in Developing
Countries. Deutsche Gesellschaft für Technische Zusammenarbeit (GTZ)
GmbH.
[28]. LASKOWSKI, R.A., MACARTHUR, M.W., MOSS, D.S., THORNTON,
J.M. (1993) J. Appl. Cryst. 26, 283-291
[29]. LÜTHY, R., BOWIE, J.U., EISENBERG, D. (1992) Nature 356, 83-85
[30]. MURZIN A. G., BRENNER S. E., HUBBARD T., CHOTHIA C. (1995) J.
Mol. Biol. 247, 536-540
[31]. MAGALHÃES, M. L. B., PEREIRA, C. P., BASSO, L. A. AND SANTOS
D. S. (2002) Prot. Exp. and Purif. 26, 59-64.
52
[32]. MICHEL, G., ROSZAK, A.W., SAUVE, V., MACLEAN, J. MATTE, A.,
COGGINS, J.R., CYGLER, M., LAPTHORN, A.J. (2003) J. Biol. Chem.
278, 19463-19472.
[33]. MILBURN, H. J. (2001) Current Opinion in Pulmonary Disease 7, 133-
141.
[34]. NORTHROP, D. B. (1975) Biochemistry 14, 2644-2651
[35]. NSB Editorial comment. Taming tuberculosis-again. Nature Structural
Biology 7, 87-88 (2000).
[36]. O´BRIEN, R. J. AND NUNN P. P. (2001) American Journal of
Respiratory and Critical Care Medicine 162, 1055-1058.
[37]. PADYANA, A.K., BURLEY, S.K. (2003) Structure 11, 1005-1013.
[38]. PABLOS-MENDEZ, A., RAVIGLIONE, M.C. et al. (1998) The New
England Journal of Medicine 338, 1641-1649.
[39]. PARISH, T. AND STOKER, N. G. (2002) Microbiology 148, 3069-3077.
[40]. PARRISH, N.M., DICK, J.D. & BISHAI, W.R. (1998) Trends in
Microbiology 6, 107-112.
[41]. PASQUALOTO, K.F.M. & FERREIRA, E.I. (2001) Current Drug Targets
2, 427-437.
[42]. PATEL, M. P., AND BLANCHARD, J. S. (2001) Biochemistry 40, 5119-
5126.
[43]. PETRINI, B. & HOFFNER, S. (1999) International Journal of
Antimicrobial Agents 13, 93-97.
[44]. QUINN, D. M., AND SUTTON, L. D. (1991) Enzyme Mechanism from
Solvent Isotope Effects (Cook, P. F., Ed.) 1st ed., pp 73-126, CRC Press,
Florida.
[45]. RATLEDGE, C. (1982) Nutrition, Growth and Metabolism in “The
Biology of the Mycobacteria” (Ratledge, C. and Stanford, J. L. Ed.) vol. 1,
pp. 185-271, Academic Press, London.
[46]. RILEY, L.W. (1993) Clinical Infectious Diseases 17, 442-446.
[47]. 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 MCLEOD,
R. (1998) Nature 393, 801-805.
[48]. RUFFINO-NETTO, A. (2002) Rev. Soc. Bras. Med. Trop. 35, 51-58.
53
[49]. STRÖMBERG, P., SVENSSON, S., BERST, K. B., PLAPP, B. V.,
HÖÖG, J. O. (2004) Biochemistry 43, 1323-1328
[50]. SINGH, S., KOROLEV, S., KOROLEVA, O., ZAREMBINSKI, T.,
COLLART, F., JOACHIMIAK, A., CHRISTENDAT, D. (2005) J. Biol.
Chem. 280, 17101-17108
[51]. SEPKOWITZ, K. A., RAFFALI, J., RILEY, L., KIEHN, T. E. AND
ARMSTRONG, D. (1995) Clinical Microbiology Reviews 180-199.
[52]. ZHANG, X., ZHANG, S., HAO, F., LAI, X., YU, H., HUANG, Y. AND
WANG, H. (2005) J Biochem Mol Biol. 38, 624-31.
[53]. Ye, S., Delft, F.V., Brooun, A., Knuth, M.W., Swanson, R.V., McRee,
D.E. (2003) J. Bacteriol. 185, 4144–4151.
[54]. YOUNG, D.B. (1998) Nature 393, 515-516.
[55]. WICKELGREN, I. (2000) Science 288, 5470, 1314-1315.
[56]. WOESTENENK, E.A., HAMMARSTRO, M., BERG, S., HÄRD, T.,
BERGLUND, H. J. (2004) Struct. Funct. Genomics 5, 217–229.
54
ANEXO I
55
Docking and small angle X-ray scattering studies
of purine nucleoside phosphorylase
Walter Filgueira de Azevedo Jr.,
a,b,
*
Giovanni C
eesar dos Santos,
a
Denis Marangoni dos Santos,
a
Johnny Rizzieri Olivieri,
a,b
Fernanda Canduri,
a,b
Rafael Guimar
~
aaes Silva,
c
Luiz Augusto Basso,
c
Gaby Renard,
c
Isabel Os
oorio da Fonseca,
c
Maria Anita Mendes,
b,d
M
aario S
eergio Palma,
b,d
and Di
oogenes Santiago Santos
c
a
Departamento de F
ıısica, UNESP, S
~
aao Jos
ee do Rio Preto, SP 15054-000, Brazil
b
Center for Applied Toxinology, Instituto Butantan, S~aao Paulo, SP 05503-900, Brazil
c
Rede Brasileira de Pesquisas em Tuberculose, Departamento de Biologia Molecular e Biotecnologia, UFRGS, Porto Alegre, RS 91501-970, Brazil
d
Laboratory of Structural Biology and Zoochemistry, Department of Biology, Institute of Biosciences, UNESP, Rio Claro, SP 13506-900, Brazil
Received 19 August 2003
Abstract
Docking simulations have been used to assess protein complexes with some success. Small angle X-ray scattering (SAXS) is a
well-established technique to investigate protein spatial configuration. This work describes the integration of geometric docking with
SAXS to investigate the quaternary structure of recombinant human purine nucleoside phosphorylase (PNP). This enzyme catalyzes
the reversible phosphorolysis of N-ribosidic bonds of purine nucleosides and deoxynucleosides. A genetic deficiency due to mu-
tations in the gene encoding for PNP causes gradual decrease in T-cell immunity. Inappropriate activation of T-cells has been
implicated in several clinically relevant human conditions such as transplant rejection, rheumatoid arthritis, lupus, and T-cell
lymphomas. PNP is therefore a target for inhibitor development aiming at T-cell immune response modulation and has been
submitted to extensive structure-based drug design. The present analysis confirms the trimeric structure observed in the crystal. The
potential application of the present procedure to other systems is discussed.
Ó 2003 Elsevier Inc. All rights reserved.
Keywords: Geometric docking; SAXS; Purine nucleoside phosphorylase; Bioinformatics
Recent developments in the algorithm for protein
docking allowed the prediction of the conformation of
quaternary structures of several proteins. Among all
available algorithms for docking of biological macro-
molecules the geometric docking has been proved to
generate reasonable models of several macromolecular
assemblies [1,2]. Small angle X-ray scattering (SAXS)
technique provides information on the structural char-
acteristics of macromolecules in solution at a super-
atomic scale. One of the procedures to obtain structural
information from SAXS results is based on the
comparison between structure functions of proposed
models with different confi gurations of monomers or
subdomains with those determined from experiments.
Even though this technique cannot guarantee the
uniqueness of the model, it is widely used and was
demonstrated to yield useful information on the struc-
ture, on structural variations, and on the quaternary
structure of a number of macromolecules of biological
interest [3]. The Guinier analysis of SAXS intensity
provides a structural parameter, the radius of gyration
of the macromolecule in solution, which is independent
of any a priori mod el. The SAXS method also yields
information on the spatial configuration of the macro-
molecular subdomains but ignores inter nal structural
details and dynamics features such as, vibration, rota-
tion or internal conformational changes [4].
Purine nucleoside phosphorylase (PNP) catalyzes the
reversible phosphorolysis of the ribonucleosides and
Biochemical and Biophysical Research Communications 309 (2003) 923–928
www.elsevier.com/locate/ybbrc
BBRC
*
Corresponding author. Fax: +55-17-221-2247.
E-mail address: [email protected] (W.F. de Azevedo Jr.).
0006-291X/$ - see front matter Ó 2003 Elsevier Inc. All rights reserved.
doi:10.1016/j.bbrc.2003.08.093
2
0
-deoxyribonucleosides of guanine, hypoxanthine, and
a number of related nucleoside compounds [5], except
adenosine. Human PNP is an attractive target for drug
design and it has been submitted to extensive structure-
based design. PNP inhibitors could be used in the fol-
lowing applications: (1) treatment of T-cell leukemia;
(2) suppression of the host versus graft response in
organ transplantation recipients; (3) treatment of sec-
ondary or xanthine gout by restricting purine catabo-
lites to the more soluble nucleosides; and (4) in
combination with nucleosides to prevent their degra-
dation by PNP metabolism [6]. More recently, the
structure of human PNP has been solved using cryo-
crystallographic techniques at 2.3
AA resolution, which
allowed a redefinition of the residues involved in the
substrate binding sites [7,8]. The crystallographic
structure is a trimer, however, there is a report of di-
meric structure for the human enzyme [9], which may
change the subunit interface. Since the active site is
located near the interface of two subunits, changing the
putative interactions between enzyme and inhibitors
should have a bearing on structure-based inhibitor
design.
Here we report the combination of geometric docking
simulations and SAXS studies to assess the human PNP
quaternary structure in solution. The general procedure,
here described, may be used to study the spatial con-
figuration of the macromolecular subdomains of other
proteins in solution.
Materials and methods
Integration of geometric docking simulation and SAXS experiments.
In order to assess the quaternary structure of PNP a scheme was used
that involved both geometric docking simulations and SAXS experi-
ments. A flowchart describing the overall strategy is shown in Fig. 1.
This procedure was used to generate the dimeric models for PNP and
the trimeric structure was built using the crystallographic symmetry. In
order to speed up the geometric docking simulations, a parallel version
of the program GRAMM [10] was used to generate the dimeric models
for human PNP. Each step of the procedure is described in the
following sections.
Geometric docking simulations. In order to obtain the dimeric
structure of human PNP, docking simulation was performed using the
geometric recognition algorithm, which was developed to identify
molecular surface complementarity. The monomeric structure of hu-
man PNP (PDB access code: 1M73) [7] was docked against its own
structure. It generated a total of 100 dimers. The geometric recognition
algorithm is based on a geometrical approach and involves an auto-
mated procedure including: (i) a digital representation of the molecules
by three-dimensional discrete functions; (ii) the calculation of a cor-
relation function that assesses the degree of molecular surface overlap
and penetration upon relative shifts of the molecules in three dimen-
sions; and (iii) a scan of the relative orientations of the molecules [10].
The procedure is equivalent to a six-dimensional search but consid-
erably faster by design, and the computation time is only moderately
dependent on molecular size. This procedure has been applied to assess
protein–protein and protein–ligand interactions. The geometric rec-
ognition algorithm was implemented in the program GRAMM [10].
All geometric docking simulations were performed on a Beowulf
cluster, with 16 nodes (B16/AMD Athlon 1800+; BioComp, S
~
aao Jos
ee
do Rio Preto, SP, Brazil).
SAXS studies. X-ray scattering data were collected at room tem-
perature using Cu Ka X-rays radiation generated by a Rigaku RU300
rotating anode generator operated at 50 kV and 90 mA and collimated
with a block slit system [11]. The scattering intensity was measured
using a linear position sensitive detector (CBPF-Brazil).
The SAXS measurements were performed within an angular range
defined by 0:02
AA
À1
< h < 0:450
AA
À1
where h ¼ð4psinÞ=k,2h being the
angle between the incident and the scattered X-ray beam and k the
X-ray wavelength. The contributions to the scattering intensity from
the solvent, capillary, and air were subtracted from the total intensity.
Recombinant human PNP was expressed and purified as previously
described [12]. The SAXS measurements were carried out using human
PNP solution, which was concentrated to 12 mg ml
À1
against 10 mM
potassium phosphate buffer (pH 7.1). The counting time was 12 h. The
extrapolated experimental SAXS intensity function was desmeared to
suppress the influence from the slit collimation system yielding the
corrected intensities, IðhÞ.
A structural parameter related to the overall size of the macro-
molecule, the radius of gyration R
g
, was determined by using the
Guinier equation [13]
IðhÞ¼Ið0Þ exp
"
À
h
2
R
2
g
3
#
: ð1Þ
Eq. (1) applies to macromolecules in the limits of a dilute solution and
small h values. More detailed information of the molecular structure
can be obtained from the distance distribution function pðrÞ, which is
related to the SAXS desmeared (free from geometrical collimation
effects) intensity IðhÞ by
Fig. 1. Flowchart describing the overall strategy to assess protein
complex conformations.
924 W.F. de Azevedo Jr. et al. / Biochemical and Biophysical Research Communications 309 (2003) 923–928
pðrÞ¼
1
2p
2
Z
1
0
IðhÞðhrÞ sinðhrÞdh: ð2Þ
The pðrÞ function is proportional to the number of pairs of electrons
separated by the distance r, which is encountered by combinations
between all the elements of the macromolecule. The radius of gyration
of macromolecules in solution is usually determined by applying Eq.
(1). The distance distribution function, p
exp
ðrÞ, has been determined by
indirect Fourier transformation using the ITP program [11]. This
program was also used to determine the intensity IðhÞ, free from
smearing collimation effects. The theoretical function p
theo
ðrÞ was cal-
culated using the program MULTIBODY [11], modified in order to
make molecular model building easier [4]. The program MULTI-
BODY calculates the resulting function pðrÞ of the complete set of
atomic coordinates of each structural model for the macromolecule. In
the present study we calculated the p
theo
ðrÞ for monomer, for the
dimers, generated by geometric docking, and for the trimer, obtained
by application of crystallographic rotations.
Correlation between geometric docking simulations and SAXS
experiments. To assess the correlation between theoretical and exper-
imental distance distribution function p
exp
ðrÞ and p
theo
ðrÞ, respectively,
we have calculated the linear correlation coefficient (CC), which is
defined as follows [14]:
CC ¼
P
n
i
jp
exp;i
ðrÞj
2
À jp
exp
ðrÞj
2

Âjp
theo;i
ðrÞj
2
À jp
theo
ðrÞj
2
hi
P
n
i
jp
exp;i
ðrÞj
2
À jp
exp
ðrÞj
2

2
P
n
i
jp
theo;i
ðrÞj
2
À jp
theo
ðrÞj
2

2

1
2
;
ð3Þ
where
jp
exp
ðrÞj
2
is the mean of the jp
exp;i
ðrÞj
2
, jp
theo
ðrÞj
2
is the mean of
jp
theo;i
ðrÞj
2
, and sums are made over all available pðr Þ. When a corre-
lation is known to be significant, CC is one conventional way of
summarizing its strength. The complex, which generates the highest
correlation coefficient, is considered the right macromolecular con-
formation. In the present work, we calculated the CC for all dimeric
models obtained from the geometric docking simulations and for the
monomeric and trimeric structures.
Results and discussion
Guinier plot (log IðhÞ versus h
2
) of the desmeared
scattering function is displayed in Fig. 2. The slope of
linear portion of this plot was determined to obtain the
radius of gyration of human PNP. The R
g
value was
29.8
AA.
PNPs from most mammalian and some of the bac-
terial sources appear to be trimeri c although dimeric
quaternary structures have been proposed for the
human enzyme [9]. Analysis of the crystallographic
structures of human PNP indicates a trimeric structure
(PDB access codes: 1ULA, 1ULB, 1M73, and 1PWY)
[6–8,15]. However, in a number of instances the qua-
ternary structure observed in the crystalline state is not
conserved in solution [3]. Furthermore, in the case of
human PNP the low pH used in the crystallization
condition [16] may generate differences in the spatial
configuration of the macromolecular subdomains ob-
served in the crystal when compared to the structure in
physiological pH. In addition, since the active site of the
PNP is located near the interface of two subunits within
the trimer, the precise information about the biological
unit in solution is of capital importance to guide the
structure-based design of inhibitors because its target is
a structure as close as possible to the structure found in
the physiological conditions, where the drug will act. Up
to now all structure-based designs of PNP inhibitors
have used the low-resolution structures of human PNP
(PDB access codes: 1ULA and 1ULB) and consider the
trimer as the target for molecular modeling studies.
Three families of theoretical models, based on the
high-resolution crystallographic structure (PDB access
code: 1M73) [7], were used to determine the theoretical
distance distribution function, p
theo
ðrÞ, using the pro-
gram MULTIBODY and then compared with the ex-
perimental function pðrÞ determined using the ITP
program from SAXS data. Figs. 3A–C show structural
models and the experimental distribution function
against the theoretical distribution function for
the monomer, dimer, and trimer, of human PNP,
respectively.
The atomic coordinates for monomeric structure
were obtained from the asymmetric unit content of the
crystallographic structure of human PNP solved at 2.3
AA
resolution [7]. Previous statistical analysis of low-reso-
lution docking indicated that gross structural features of
protein–protein interactions could be identified for a
significant percentage of protein complexes [1]. There-
fore, the low-resolution protocol of the GRAMM pro-
gram [10] was used to generate the dimeric models. A
total of 100 models for the dimeric structure were built,
only the complex, which generated the highest correla-
tion coefficient between theoretical and experimental
distance distribution function is shown in Fig. 3B. The
trimeric structure was built applying two successive
rotations of 120° along z-axis on the atomic coordinat es
of the monomer. The radii of gyration for the structural
models are 18.7, 26.5, and 28.7
AA, for the monomer,
dimer, and trimer, respectively.
Fig. 2. Guinier plot of the SAXS intensities, IðhÞ, for human PNP. The
straight line was obtained by least-squares fitting in the region
h
2
< 1:8:10
À3
AA
2
.
W.F. de Azevedo Jr. et al. / Biochemical and Biophysical Research Communications 309 (2003) 923–928 925
The value CC lies betw een )1 and 1. It has a value of
1, when the data points lie on a perfect straight line with
positive slope. If the data points lie on a perfect straight
line with negative slope, then CC has the value )1 [14].
The correlation coefficient between theoretical and ex-
perimental distance distribution functions ranges from
0.591 to 0.995, an d the highest correlati on coefficient
was obtained for the trimeric structure, which also
presented the radius of gyration closer to the experi-
mental radius of gyration.
The contact area at interface between each subunit in
the PNP trimer is 1124
AA
2
, which indicates that the
subunits are strongly bound to each other. Fig. 4 shows
the electrostatic potential surface at subunit interface of
the trimeric structure generat ed using GRASP [17].
Analysis of the electrostatic potential surface at the
Fig. 3. Proposed structural models for human PNP and the corresponding theoretical (continuous line) and experimental (dotted line) distance
distribution functions, pðrÞ, for: (A) monomer, (B) dimer, and (C) trimer. The model figures were generated by MOLSCRIPT [27] and Raster3D [28].
926 W.F. de Azevedo Jr. et al. / Biochemical and Biophysical Research Communications 309 (2003) 923–928
subunit interface indicates good shape complementarity
and some charge complementarity; however, most of the
contacts are hydrophobic and involve residues Tyr88,
Phe141, Phe159, Phe200, and Leu209.
The trimeric PNP struc ture has been extensively used
for structure-based studies of PNP inhibitors [5,6,18–
25]. However, the quaternary structure of human PNP
in solution and in physiological pH has not been pre-
viously investigated using low-resolution methods, such
as SAXS. The present analysis of the SAX S experiments
integrated with geomet ric docking simulation strongly
indicates that human PNP is a trimer in solution, the
agreement found between the experimental and theo-
retical pðrÞ functions for the trimer suggests that struc-
ture in solution adopts approximately the same
conformation identified in the high-resolution crystal-
lographic structure (PDB access code: 1M73) [7]. The
radius of gyration determined for the trimeric structure
is slightly smaller than that determined from the Guinier
plot (log IðhÞ versus h
2
) of the desmeared scattering
function. The possible reasons for this discrepancy may
be the cryogenic conditions used to solve the high-res-
olution structure of human PNP and the absence of
solvents in the theoretical model.
The integration of a high-efficient algorithm for
geometric docking with SAXS experiments allowed the
investigation of the possible quaternary structures not
observed in the crystalline state, such as the putative
PNP dimeric structure [9]. The pr ocedure adopted to
analyze the interaction between PNP subunits can be
used for other protein complexes. The main applications
of the present methodology are: (1) analysis of interac-
tions between biological macromolecules using struc-
tural models obtained from crystallography or NMR,
(2) validation of structural models obtained from
molecular modeling [26] of complexes of biological
macromolecules, and (3) analysis of complexes of
biological macromolecules in conditions closer to the
biological environment.
Geometric docking simulations may be omitted from
the strategy if the atomic coordinates for the complexes
are available. We are applying the procedure, here de-
scribed, to assess the quaternary structure of a number
of protein complexes, such as hemoglobins, PNPs, and
crotoxin.
Acknowledgments
This work was supported by grants from FAPESP (SMOLBNet,
Proc. Num. 01/07532-0), CNPq, CAPES and Instituto do Millenium
(CNPq-MCT). W.F.A. (CNPq, 300851/98-7), M.S.P. (CNPq, 500079/
90-0), and L.A.B. (CNPq, 520182/99-5) are researchers for the Bra-
zilian Council for Scientific and Technological Development.
References
[1] A. Tovchigrechko, C.A. Wells, I.A. Vakser, Docking of protein
models, Protein Sci. 11 (2002) 1888–1896.
[2] N. Jing, C. Marchand, J. Liu, R. Mitra, M.E. Hogan, Y.
Pommier, Mechanism of inhibition of HIV-1 integrase by G-
tetrad-forming oligonucleotides in vitro, J. Biol. Chem. 275 (2000)
21460–21467.
[3] D.I. Svergun, C. Barberato, M.H.J. Koch, L. Fetler, P. Vachette,
Large differences are observed between the crystal and solution
quaternary structures of allosteric aspartate transcarbamylase in
the R state, Proteins Struct. Funct. Genet. 27 (1997) 110–117.
[4] J.R. Olivieri, A.F. Craievich, The subdomain structure of human
serum albumin in solution under different pH conditions studied
by small angle X-ray scattering, Eur. Biophys. J. 24 (1995) 77–
84.
[5] J.A. Montgomery, Purine nucleoside phosphorylase: a target for
drug design, Med. Res. Rev. 13 (1993) 209–228.
[6] S.E. Ealick, Y.S. Babu, C.E. Bugg, M.D. Erion, W.C. Guida, J.A.
Montgomery, J.A. Secrist III, Application of crystallographic and
modeling methods in the design of purine nucleoside phosphor-
ylase inhibitors, Proc. Natl. Acad. Sci. USA 91 (1991) 11540–
11544.
Fig. 4. Electrostatic potential surface at subunit interface of human PNP, calculated with GRASP [17] and shown from )10 to +10 kT. Uncharged
regions are white.
W.F. de Azevedo Jr. et al. / Biochemical and Biophysical Research Communications 309 (2003) 923–928 927
[7] W.F. de Azevedo Jr., F. Canduri, D.M. Santos, R.G. Silva, J.S.
Oliveira, L.P.S. Carvalho, L.A. Basso, M.A. Mendes, M.S. Palma,
D.S. Santos, Crystal structure of human purine nucleoside
phosphorylase at 2.3
AA resolution, Biochem. Biophys. Res. Com-
mun. 308 (2003) 545–552.
[8] D.M. Santos, F. Canduri, J.H. Pereira, M.V.B. Dias, R.G. Silva,
M.A. Mendes, M.S. Palma, L.A. Basso, W.F. de Azevedo, D.S.
Santos, Crystal structure of human purine nucleoside phosphor-
ylase complexed with acyclovir, Biochem. Biophys. Res. Com-
mun. 308 (2003) 553–559.
[9] A.S. Lewis, B.A. Lowy, Human erythrocytes purine nucleoside
phosphorylase: molecular weight and physical properties, J. Biol.
Chem. 254 (1979) 9927–9932.
[10] E. Katchalski-Katzir, I. Shariv, M. Eisenstein, A.A. Friesem, C.
Aflalo, I.A. Vakser, Molecular surface recognition: determination
of geometric fit between proteins and their ligands by correlation
techniques, Proc. Natl. Acad. Sci. USA 89 (1992) 2195–2199.
[11] O. Glatter, in: O. Glatter, O. Kratky (Eds.), Small Angle X-Ray
Scattering, Academic Press, London, 1982.
[12] R.G. Silva, L.P. Carvalho, J.S. Oliveira, C.A. Pinto, M.A.
Mendes, M.S. Palma, L.A. Basso, D.S. Santos, Cloning, overex-
pression, and purification of functional human purine nucleoside
phosphorylase, Protein Expr. Purif. 27 (2003) 158–164.
[13] A. Guinier, G. Fournet, Small-Angle Scattering of X-rays, Wiley,
New York, 1955.
[14] W.H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannery,
Numerical Recipes in FORTRAN. The Art of Scientific Com-
puting, second ed., Cambridge University Press, New York, 1992.
[15] S.E. Ealick, S.A. Rule, D.C. Carter, T.J. Greenhough, V. Babu,
W.J. Cook, J. Habash, J.R. Helliwell, J.D. Stoeckler, R.E. Parks
Jr., F. Chen, C.E. Bugg, Three-dimensional structure of human
erythrocytic purine nucleoside phosphorylase at 3.2
AA resolution,
J. Biol. Chem. 265 (1990) 1812–1820.
[16] W.J. Cook, S.E. Ealick, C.E. Bugg, J.D. Stoeckler, R.E. Parks Jr.,
Crystallization and preliminary X-ray investigation of human
erythrocytic purine nucleoside phosphorylase, J. Biol. Chem. 256
(1981) 4079–4080.
[17] A. Nicholls, K. Sharp, B. Honig, Protein folding and association:
insights from the interfacial and thermodynamic properties of
hydrocarbons, Proteins Struct. Funct. Genet. 11 (1991) 281–296.
[18] P.W.K. Woo, C.R. Kostlan, J.C. Sircar, M.K. Dong, R.B.
Gilbertsen, Inhibitors of human purine nucleoside phosphorylase.
Synthesis and biological activities of 8-amino-3-benzylhypoxan-
thine and related analogues, J. Med. Chem. 35 (1992) 1451–1457.
[19] J.-W. Chern, H.-Y. Lee, C.-S. Chen, Nucleosides. 5. Synthesis of
guanine and formycin B derivatives as potential inhibitors of
purine nucleoside phosphorylase, J. Med. Chem. 36 (1993) 1024–
1031.
[20] M.D. Erion, S. Niwas, J.D. Rose, S. Ananthan, M. Allen, J.A.
Secrist III, Y.S. Babu, C.E. Bugg, W.C. Guida, S.E. Ealick, J.A.
Montgomery, Structure-based design of inhibitors of purine
nucleoside phosphorylase. 3. 9-Arylmethyl derivatives of 9-
deazaguanine substituted on the methylene group, J. Med. Chem.
36 (1993) 3771–3783.
[21] J.A. Secrist III, S. Niwas, J.D. Rose, Y.S. Babu, C.E. Bugg, M.D.
Erion, W.C. Guida, S.E. Ealick, J.A. Montgomery, Structure-
based design of inhibitors of purine nucleoside phosphorylase. 2.
9-Alicyclic and 9-heteroalicyclic derivatives of 9-deazaguanine, J.
Med. Chem. 36 (1993) 1847–1854.
[22] W.C. Guida, R.D. Elliot, H.J. Thomas, J.A. Secrist III, Y.S.
Babu, C.E. Bugg, M.D. Erion, S.E. Ealick, J.A. Montgomery,
Structure-based design of inhibitors of purine nucleoside phos-
phorylase. 4. A study of phosphate mimics, J. Med. Chem. 37
(1994) 1109–1114.
[23] S. Niwas, P. Chand, V.P. Pathak, J.A. Montgomery, Structure-
based design of inhibitors of purine nucleoside phosphorylase. 5.
9-Deazahypoxanthines, J. Med. Chem. 37 (1994) 2477–2480.
[24] P.E. Morris, A.J. Elliott, S.P. Walton, C.H. Williams, J.A.
Montgomery, Synthesis and biological activity of a novel class of
purine nucleoside phosphorylase inhibitors, Nucleosides Nucleo-
tides Nucleic Acids 19 (2000) 379–404.
[25] A. Andricopulo, R.A. Yunes, Structure–activity relationships for
a collection of structurally diverse inhibitors of purine nucleoside
phosphorylase, Chem. Pharm. Bull. 49 (2001) 10–17.
[26] A. Sali, T.L. Blundell, Comparative protein modelling by satis-
faction of spatial restraints, J. Mol. Biol. 234 (1993) 779–815.
[27] P.J. Kraulis, MOLSCRIPT: a program to produce both detailed
and schematic plots of proteins, J. Appl. Cryst. 24 (1991) 946–950.
[28] E.A. Merritt, D.J. Bacon, Raster3D: photorealistic molecular
graphics, Methods Enzimol. 277 (1997) 505–524.
928 W.F. de Azevedo Jr. et al. / Biochemical and Biophysical Research Communications 309 (2003) 923–928
ANEXO II
56
Selection of an Escherichia coli host that expresses mutant forms
of Mycobacterium tuberculosis 2-trans enoyl-ACP(CoA) reductase
and 3-ketoacyl-ACP(CoA) reductase enzymes
Simone S. Poletto,
a
Isabel O. da Fonseca,
a
Luiz P.S. de Carvalho,
a
Luiz A. Basso,
a,
*
and Di
ogenes S. Santos
b,
*
a
Departamento de Biologia Molecular e Biotecnologia, Universidade Federal do Rio Grande do Sul, Avenida Bento Gonc
ß
alves, 9500,
Porto Alegre, RS 91501-970, Brazil
b
Faculdade de Farmacia Pontif
ıcia, Universidade Catolica do Rio Grande do Sul, Porto Alegre, RS 90619-900, Brazil
Received 14 August 2003, and in revised form 17 October 2003
Abstract
Tuberculosis (TB) still remains a worldwide health concern. Efforts to understand the complex biology of Mycobacterium
tuberculosis, the causative agent of TB, are important for new antitubercular drug development. Despite the completion of the
genome sequence and the development of new genetic tools to manipulate this organism, the availability of sufficient amounts
of mycobacterial proteins still remains an essential and laborious step to study the biochemical features of this pathogen. The
T7-RNA polymerase-based pET system has been largely employed to express mycobacterial proteins in Escherichia coli, but it
presents some limitations. To overcome problems with unstable expression of an M. tuberculosis inhA-encoded enoyl reductase
mutant protein and lack of expression of two mabA-encoded ketoacyl reductase mutants, a sub-population of E. coli
BL21(DE3) host cells was selected from a small-opaque colony. This empirically selected host, named BL21(DE3)NH, allowed
stable expression of these mutant proteins. Although the mechanism that led the BL21(DE3)NH host to express the re-
combinant mutant proteins remains unknown, the persistent phenotype points to a stable genetic switch. This genetic alteration
resulted in a tight control of the highly processive T7 RNA polymerase. Moreover, the absolute requirement for IPTG to
obtain protein expression in the BL21(DE3)NH host cells suggests that no inherent defect in the transcriptional activity of the
T7 promoter is present. Empirical host selection requires no further genetic manipulation of recombinant plasmids and may
represent a means of obtaining tailor-made E. coli strains that overcome toxic effects associated with heterologous protein
expression.
Ó 2003 Elsevier Inc. All rights reserved.
Keywords: Protein expression; Escherichia coli; Host selection; Mycobacteria
Tuberculosis (TB), caused by Mycobacterium tuber-
culosis, remains one of the deadliest diseases in the
world. It is estimated that 8.2 million new TB cases
occurred worldwide in the year 2000, 1.8 million deaths
occurred in the same year, and more than 95% of those
were in developing countries [1]. Possible factors un-
derlying the resurgence of TB include the HIV epidemic,
increase in the homeless population, and decline in
health care structures and national surveillance [2].
Another contributing factor is the evolution of multi-
drug-resistant strains of M. tuber culosis (MDR-TB),
defined as resistant to at least isoniazid and rifampicin,
which are the most effective first-line drugs [3]. Treat-
ment of MDR-TB requires the administration of sec-
ond-line drugs (amikacin, kanamycin, capreomycin,
cycloserine, para-aminosalicylic acid, ethionamide, and
fluoroquinolones) that are more toxic and less effective
and are given for at least three times as long and at 100
times the cost of basic short-course chemotherapy regi-
mens [4]. Hence, new antimycobacterial agents are need-
ed to improve the treatment of MDR-TB, to shorten the
treatment course to improve patient compliance, and to
*
Corresponding author. Fax: +55-51-33167309.
E-mail addresses: [email protected] (L.A. Basso),
[email protected] (D.S. Santos).
1046-5928/$ - see front matter Ó 2003 Elsevier Inc. All rights reserved.
doi:10.1016/j.pep.2003.10.009
Protein Expression and Purification 34 (2004) 118–125
www.elsevier.com/locate/yprep
provide for more effective treatment of latent tubercu-
losis infection [5].
The study of biology, pharmacology, and host–
pathogen interactions is important for the development
of new drugs as well as vaccines against M. tuber culosis.
The completion of the genome sequence of M. tuber-
culosis [6] and availability of genetic tools to manipulate
this organism [7] have accelerated these studies. Despite
these advances, the availability of sufficient amounts of
proteins of M. tuberculosis still remains an essential and
laborious step to study the biochemical features of this
pathogen. This fact is due to (i) the slow-growth of the
bacillus whose generation time is 24 h; (ii) since aerosols
containing bacteria are the way of spread of the disease,
a P3 facility is required to grow the bacillus culture; (iii)
and the highly lipophilic envelope of mycoba cteria that
represents an obstacle to protein extraction and purifi-
cation. On the other hand, the improvements in the re-
combinant DNA technology, the advances in molecular
genetic studies in mycobacteria and, as previously
pointed out, the availability of the complete genome
sequence have facilitated the cloning and he terologous
expression of mycobacterial proteins. Escherichia coli
1
and Mycobacterium smegmatis (a fast-growing bacte-
rium that belongs to the same genus of M. tuberculosis)
are the first choices for hosts [8]. E. coli, in particular,
remains one of the most attractive systems for heterol-
ogous protein expression due to its ability to grow
rapidly and at high density on inexpensive substrates, its
well-characterized genetics, and the avail ability of an
increasingly large number of cloning vectors and mutant
host strains [9]. The pET system (Novagen) developed
by Studier and Moffat [10] has been one of the most
widely used systems for expression of mycobacterial
proteins in E. coli [11–15]. In this system, target genes
are positioned downstream of the bacteriophage T7 late
promoter on medium copy number plasmids. The highly
processive T7 RNA polymerase gene (DE3 lysogen) of
production hosts is placed under control of the IPTG-
inducible lacUV5 promoter. However, expression of
membrane proteins and proteins toxic to the host cells
[16], formation of inclusion bodi es [17] or complete
absence of expression are amongst the difficulties en-
countered with mycobacterial proteins.
The inhA-encoded 2-trans enoyl-ACP (CoA) reduc-
tase (ENR) and mabA-encoded 3-ketoacyl-ACP (CoA)
reductase (KAR) are enzymatic components of the
Type II fatty acid synthase (FAS-II) of mycobacteria
and are thus involved in mycolic acid biosynthesis [18].
The mycobacterial NADH-dependent ENR is the tar-
get for isoniazid [11,19], the most prescribed drug to
treat tuberculosis, making it an excellent target for TB
drug development. Another reductive step of the
mycobacterial FAS-II system is catalyzed by the
NADPH-dependent KAR enzyme [20]. The FAS-II
system enzymes have been shown to be essential for
mycobacterial growth and configured into high
throughput screening for identification of new drugs
against TB [21] . The amino acid residues tyrosine-158
and lysine-165 have been implicated in the catalytic
mechanism of M. tuberculosis ENR [11,22,23]. More-
over, the side chain of Y158 of M. tuberculosis ENR
has been shown to unde rgo a rotation upon binding of
triclosan, an e nzyme inhibi tor, to form a hydrogen
bond with the hydroxyl group of this inhibitor [18].
The mycobacterial KAR and ENR enzymes are
members of the short-chain dehydrogenase/reductase
(SDR) family and thereby display the amino acid sig-
nature of this family, S–(X)
12
–Y–(X)
3
–K [24]. As part
of our efforts to develop novel antitubercular drugs
based on rational design, we are studying the catalytic
mechanism of M. tuberculosis ENR and KAR en-
zymes, identifying amino acid residues involved in ei-
ther catalysis or substrate binding, and trying to solve
their three-dimensional structures. Accordingly, seven
mutant proteins were produced for the putative cata-
lytic amino acids (ENR: Y158F, K165A, K165Q;
KAR: S140T, S140A, Y153F, and K157A), cloned into
the pET system, and transformed into E. coli
BL21(DE3) host strain. However, the Y158F ENR
mutant failed to show stable protein expression in this
system, and no expression could be obtained for the
Y153F and K157A KAR mutant enzymes.
Miroux and Walker [16] have selected mutant hosts
derived from E. coli BL21(DE3) that allowed protein
expression at higher levels than the parent host. The
mutant host C41(DE3) has been employed to express E.
coli inner membrane enzyme acyl acyl carrier protein
synthase [25] and M. tuberculosis acyl carrier protein
(AcpM) and malonyl-CoA:AcpM transacylase [15].
Building on the strategy adopted by Miroux and Walker
[16], we selected a sub-population of E. coli BL21(DE3),
named BL21(DE3)NH, that was able to express
theY158F ENR, Y153F KAR, and K157A KAR mu-
tant proteins, which could not be expressed in the E. coli
BL21(DE3) parent strain. Here, we report site-directed
mutagenesis to obtain theY158F ENR, Y153F KAR,
and K157A KAR mutants, and selection of a host cell
that was able to express these mutant proteins in soluble
form. Scanning electron microscopy has been performed
to evaluate changes in BL21(DE3)NH morphology as
compared to the BL21(DE3) parent strain.
1
Abbreviations used: E. coli, Escherichia coli; ENR, enoyl reduc-
tase; InhA, enoyl-ACP (CoA) reductase from Mycobacterium tuber-
culosis; KAR, 3-ketoacyl-ACP (CoA) reductase; LB, Luria–Bertani
medium; SDS–PAGE, sodium dodecyl sulfate–polyacrylamide gel
electrophoresis; IPTG, isopropyl-b-
DD
-thiogalactopyranoside; MabA,
3-ketoacyl-ACP (CoA) reductase from Mycobacterium tuberculosis;
PCR, polymerase chain reaction; bp, base pairs; WT, wild type; EtdBr,
ethidium bromide; SEM, scanning electron microscopy; OsO
4
, osmium
tetraoxide; expression plasmid names consist of the name of the parent
plasmid vector followed by the recombinant DNA insert they encode.
S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125 119
Materials and methods
Materials
The pET-23d(+)::inhA recombinant plasmid was a
gift from Dr. John S. Blanchard at Albert Einstein
College of Medicine (AECOM, Bronx, NY) and pET-
3d::mabA recombinant plasmid was a gift from Dr.
William R. Jacobs Jr. at AECOM (Bronx, NY). pET-
23d(+), pET- 23a(+), BL21(DE 3) E. coli strain, and
Perfect Protein Marker were purchased from Novagen;
Luria–Bertani medium (LB), oligonucleotides, T4 DNA
ligase, IPTG, DH10B cells, and DNA and protein mo-
lecular weight standards were from Gibco-BRL; car-
benicillin was from Sigma; BamHI, HindIII, NcoI, and
NdeI restriction enzymes were from Boehringer–Mann-
heim and New England Biolabs; Pfu DNA polymerase
was from Stratagene; Agarose was from SEA KEM
GTG; Zero Blunt PCR Cloning Kit pCR-Blunt and
Top10 One Shot were from Invitrogen; Quantum Prep
Plasmid Miniprep kit, QIAEX II Agarose Gel Extrac-
tion kit, and QIAprep Spin Miniprep kit were from
Qiagen; Thermo Sequenase radiolabeled terminator cy-
cle sequencing kit was from Amersham Life Science; and
SDS–polyacrylamide gels and Tris–glycine running
buffer were from Bio-Rad.
Cloning of wild-type mabA gene into pET23a(+) vector
The synthetic oligonucleotide primers used in this
study are listed in Table 1. The primers mabA5, con-
taining a NdeI restriction site (bold), and mabA3, con-
taining a HindIII restriction site (bold), were used to
amplify the mabA gene from the pET-3d:: mabA re-
combinant plasmid using Pfu DNA polymerase at
standard conditions. The PCR product was purified
from agarose gel using QIAEX II Kit and inserted into a
pCR-blunt vector. E. coli Top 10 One-Shot competent
cells were transformed with the recombinant plasmid.
Plasmid DNA was extracted using Quantum Prep
Plasmid Miniprep Kit (Qiagen). The insert was removed
by NdeI and HindIII double digestion and ligated into a
pET23a(+) digested with the same restriction enzymes.
The integrity of the mabA gene was confirmed by DNA
sequencing.
Site-directed mutagenesis of inhA an d mabA genes
Site-directed mutations were introduced in the inhA
and mabA genes using the method described by Ho et al.
[26]. A list of the primers used for all amplification steps
is given in Table 1. It should be noted that only the
primers to obtain theY158F ENR, Y153F KAR, and
K157A KAR proteins by site-directed mutagenesis are
described here, since these were the mutants that showed
unstable or no protein expression. The experimental
protocol described here is exemplified by Y158F inhA
mutant gene production. First, two independent PCRs
using Pfu DNA polymerase were performed using
primers P1 and Y158b
reaction 1, and JSB02 and
Y158a
reaction 2, and the plasmid pET23d (+)::inhAas
template. The PCR program was as follows: one step of
98 °C for 3 min, 35 cycles at 98 °C for 1 min, 65 °C for
1 min, and 72 °C for 2 min, followed by a final extension
step at 72 °C for 2 min, using a DNA Thermal Cycler
(PTC200, MJ, USA). The PCR products of these reac-
tions were purified by electrophoresis on 1% agarose gel
and DNA was extracted using the QIAEX II Kit. A
second PCR was performed using the products of the
reactions 1 and 2 as templates and the primers P1 and
JSB02. The overlap PCR conditions were identical to
those used in the first round of amplification. The PCR-
amplified band was purified and cloned as described in
the previous section, yielding pET23d(+)::inhAY158F
recombinant plasmid. The primers used to make the
mabA mutants are listed in Table 1 and followed the
same protocol just described, yielding pET23a(+)::ma-
bAY153F and pET23a(+)::mabAK157A recombinant
plasmids. The sequence of the DNA inserts in all re-
combinant plasmids was confirmed by the dideoxynu-
cleotide chain termination method using the Thermo
Sequenase radiolabeled terminator cycle sequencing kit
(Amersham Biosciences).
Protein expression of WT and ENR mutants in
BL21(DE3)
To assess the expression of these constructs,
BL21(DE3) electrocompetent cells were transformed
with the following recombinant plasmids: pET23-
d(+)::inhA and pET23d(+)::inhAY158F and grown in
LB medium containing carbenicillin (50 lgml
À1
)at
various times after induction with IPTG (1 mM) at
37 °C. Control experiments were performed under the
same experimental conditions, except that E. coli
Table 1
Primers used for amplification and mutagenesis of inhA and mabA
genes from M. tuberculosis H37Rv
Primers Restriction
sites
Sequence
P1 NcoI5
0
-att gaa cca tgg cag gac tgc tgc acg-3
0
JSB02 BamHI 5
0
-gcggatccgctagagcaattgggtgtgcgc-3
0
Y158a 5
0
-atg ccg gcc ttc aac tgg atg-3
0
Y158b 5
0
-cat cca gtt gaa ggc cgg ca t-3
0
mabA5 NdeI5
0
-att cat atg act gcc aca gcc act gaa gg-3
0
mabA3 HindIII 5
0
-t aag ctt tca gtg gcc cat acc cat gcc-3
0
Y153F5 5
0
-cag gcc aac ttc gca gcc tcc-3
0
Y153F3 5
0
-gga ggc tgc gaa gtt ggc ctg-3
0
K157A5 5
0
-gca gcc tcc gcg gcc gga gtg-3
0
K157A3 5
0
-cac tcc ggc cgc gga ggc tgc-3
0
The sequences underlined correspond to degenerate codons and the
sequences in bold correspond to restriction sites.
120 S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125
BL21(DE3) host cells harbored the expression vector
lacking the target gene. Cells were harvested (14,000g
for 10 min) at 4, 8, 12, 16, 20, and 24 h after induction,
lysed by sonication, and cell debris was removed by
centrifugation (20,000g for 20 min). The resulting solu-
ble protein content was analyzed by 12% SDS–PAGE.
Stable host production
Escherichia coli BL21(DE3) electrocompetent cells
were transformed with pET23d(+)::inhAY158F, plated
on LB solid medium containing carbenicillin
(50 lgml
À1
), and grown overnight at 37 °C, yielding sub-
populations of large-transluc ent and small-opaque col-
onies. These two types of colonies were examined for
their ability to express Y158F ENR protein in liquid
culture, using the protocol of IPTG induction as previ-
ously described for wild-type ENR. The degree of pro-
tein expression, as monitored by SDS–PAGE, was
dependent on colony morphology, with smaller colonies
expressing the desired protein, whereas larger colonies
did not. Since sub-culturin g may eventually result in loss
of phenotype, E. coli BL21(DE 3) host cells wer e trans-
formed with pET23d(+)::inhAY158F recombinant plas-
mid and plated on LB–carbenicillin overnight at 37 °C. A
single small colony was used to inoculate 5 ml LB–car-
benicillin medium, grown to midlog phase (0.4–0.6
OD
600
)at37° C, and induced with 1 mM IPTG for 3 h.
At midlog phase of growth, an aliquot was withdrawn to
prepare a glycerol culture stock. The remaining of cells
were harvested by centrifugation at 14,000g for 10 min.
Pelleted material was solubilized in 400 llof20mM
Pipes buffer, pH 7.3, disrupted by sonication, and the
resulting crude extract was centrifuged at 14,000g for
10 min. Expression of the recombinant protein in soluble
form was confirmed by 12% SDS–PAGE. The glycerol
stock of this recombinant protein expressing host was
cured of pET-23d(+)::inhAY158F by growth in LB li-
quid medium in the absence of carbenicillin for 15 days.
Each day, a portion of the culture (5 ll) was used to in-
oculate 5 ml of LB. After 15 days, small-colony derived
BL21(DE3) cells lacking the recombinant plasmid arose,
which was verified both by its inability to grow on LB
solid medium containing carbenicillin and no recovery of
recombinant plasmid from the cells. To confirm the
ability of the new host, henceforth named
BL21(DE3)NH (NH standing for New Host), to confer
stable overexpression of Y158F ENR protei n,
BL21(DE3)NH electrocompetent cells were transformed
with either pET-23d(+) plasmid as control or pET-
23d(+)::inhAY158F recombinant plasmid, plated on LB-
agar plate containing carbenicillin (50 lgml
À1
), and
grown overnight at 37 °C. Seven single colonies were
separately used to inoculate 5 ml LB–carbenicillin liquid
medium, grown to an OD
600
of 0.4–0.6, induced with
1 mM IPTG, and grown for 3 h at 37 °C after induction.
The cells were harvested by centrifugation at 14,000g for
10 min, pelleted material was solubilized in 400 llof
20 mM Pipes buffer, pH 7.3, and disrupted by sonication.
Soluble proteins were analyzed by SDS–PAGE.
Protein expression of mycobacterial KAR mutants in
BL21(DE3)NH host cells
To address the ability of this new host to express pro-
teins that, in our hands, have failed to do so in the
BL21(DE3) parent host, BL21(DE 3), and BL21(DE3)
NH electrocompetent cells were transformed with the
following plasmids: pET-23a(+), pET-23a(+)::mabAY
153F and pET-23a(+)::mabAK157A, plated on LB–car-
benicillin medium, and incubated overnight at 37 °C.
Single colonies were used to inoculate 5 ml LB medium
containing carbenicillin (50 lgml
À1
) and 1 mM of IPTG
was added to the liquid cultures as they reached an OD
600
between 0.4 and 0.6. The cultures were grown at 37 °C for
8 h as determined for the wild-type KAR protein in
BL21(DE3) (data not shown). Cells were harvested by
centrifugation (14,000g for 10 min) and pellets were sol-
ubilized in 400 ll of 50 mM Tris–HCl, pH 7.8, buffer,
disrupted by sonication, and clarified by centrifugation
(14,000g for 10 min). The soluble fraction of the cellular
extracts was analyzed by SDS–PAGE.
Scanning electronic microscopy
The procedure for scanning electronic microscopy of
E. coli was carried out as described elsewhere [27,28].
Slides of E. coli BL21(DE3) and BL21(DE3)NH were
obtained by cutting single colonies (1 mm
2
and 0.5 mm of
thickness) from the LB agar plates, followed by a 2-h
immersion in Karmovisk solution at room temperature to
chemically stabilize the sample and an additional fixation
with 1% OsO
4
. Samples were then washed 10 times with
cacodylate buffer solution. For dehydration, samples
were washed with 30–100% ethanol and, subsequently,
with aqueous carbonic acid. The specimens were coated
with gold by the sputtering process to obtain a good sec-
ondary electron production for image formation. Mi-
crographs were taken using Zeiss DMS 940A and
operated in electron emission mode at 20 kV accelerating
voltage. Samples were observed in 3000Â amplification at
12 mm distance.
Results and discussion
Site-directed mutagenesis of both inhAandmabA
genes was performed by the overlap extension method
[26] and mutations were confirmed by DNA sequencing.
The inhA mutant was cloned into pET-23d(+) and
mabA mutants were inserted into pET-23a(+) expres-
sion plasmid.
S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125 121
Overexpression of mycobacterial K165A, K165Q,
and Y158F ENR mutant protei ns is shown in Fig. 1.
However, mycobacterial Y158F ENR mutant protein
overexpression was not stable, that is, not all colonies
picked up at random expressed the recombinant protein.
The appearance of two types of colonies with distinct
morphologies (large and small) of BL21(DE3) trans-
formed with pET-23d(+)::inhAY158F plasmid was ob-
served on LB–carbenicillin plates incubated overnight at
37 °C. The small colonies, which have diameters ap-
proximately 30% smaller than those of large colonies,
were opaque, while the large colonies were translucent.
In addition, the small-opaque colonies expressed the
desired proteins while the large-translucent colonies did
not. Variations in colony morphology of E. coli hosts
have been associated with detection of recombinant
colonies [29] and expression of the recombinant gene
[30]. Moreover, empirical selection has been employed
by Miroux and Walker [16] to tailor expression hosts to
overcome toxic effects associated with protein overex-
pression. Building on Miroux an d Walker strategy, a
sub-population of BL21(DE3) cells having a small-
opaque phenotype that expressed the Y158F ENR
mutant protein was selected, isolated, and cured. The
selected host cell, named BL21(DE3)NH, transformed
with pET-23d(+)::inhAY158F and plated on LB–car-
benicillin solid medium, yielded a population of small
colonies only. To evaluate whether protein expression of
the recombinant protein requires IPTG induction, seven
isolated colonies were picked up at random, suspended
in 200 ll LB–carbenicillin medium and 50 ll of the re-
sulting suspension used to inoculate 5 ml LB–carbeni-
cillin medium, and the experimental protocol described
in Material and methods was followed. Fig. 2A shows a
representative result for a single colony, suggesting that
expression of Y158 ENR in the BL21(DE3)NH host can
be achieved only in the presence of the inducer. Fig. 2B
shows that all seven colonies express the recombinant
Y158F ENR mutant protein in soluble form upon IPTG
induction.
A protocol had been designed to optimize protein
expression of mycobacterial wild-type KAR, involving
addition of 1 mM IPTG to an LB–carbenicillin liquid
culture (0.4 < OD
600 nm
< 0.6) of BL21(DE3) and further
growth for 8 h at 37 °C. The mycobacterial Y153F and
K157A KAR mutants, however, did not show any ex-
pression under these experimental conditions (Fig. 3A).
To test the ability of the empirically selected host cells to
express these mutants, BL21(DE3)NH cells were trans-
formed with pET23a(+)::mabAY153F and pE-
T23a(+)::mabAK157A recombinant plasmids, and
pET23a(+) plasmid (control). LB–carbenicill in liquid
medium was inoculated with single colonies, and cell
growth and induction was carried out as for the
Fig. 1. SDS–PAGE (12%) of the soluble fraction of the cell extracts
expressing M. tuberculosis ENR proteins upon IPTG induction in E.
coli BL21(DE3) host cells. Lane M, Protein molecular weight stan-
dards (Gibco-BRL: insulin a and b chains, 3 kDa; bovine trypsin in-
hibitor, 6.2 kDa; lysozyme, 14.3 kDa; b-lactoglobulin, 18.4 kDa;
carbonic anhydrase, 29 kDa; and ovalbumin, 43 kDa); lane 1, wild-
type ENR; lane 2, pET-23d(+); lane 3, K165A ENR; lane 4, K165Q
ENR; and lane 5, Y158F ENR. The arrow points the band corre-
sponding to M. tuberculosis ENR proteins ($29 kDa).
Fig. 2. SDS–PAGE (12%) of the soluble fraction of the cell extracts of BL21(DE3)NH transformed with either pET-23d(+) (control) or pET-
23d(+)::inhAY158F recombinant plasmid. (A) Lane 1, 1 mM IPTG; lane 2, no IPTG; lane 3, control with 1 mM IPTG; lane 4, control in the absence
of IPTG; and lane M, protein molecular weight standards (same as Fig. 1). (B) Lane 1, control with 1 mM IPTG; lane M, protein molecular weight
standards (same as Fig. 1); and lanes 2–8, protein expression of seven colonies of BL21(DE3)NH transformed with pET-23d(+)::inhAY158F plasmid
upon 1 mM IPTG induction. The arrows point the band corresponding to mycobacterial ENR ($29 kDa).
122 S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125
wild-type KAR enzyme. Cells were harvested and the
resulting pellets were suspend ed in Tris–HCl 50 mM, pH
7.8, disrupted by sonication and the solubl e fractions
were analyzed by SDS–PAGE. High level of protein
expression was observed in BL21(DE3)NH host cells
only upon IPTG induction (Fig. 3B).
The intrinsic toxicity to E. coli BL21(DE3) host cells of
pET vectors harboring no insert in the presence of IPTG
[16] has not been observed for the BL21(DE3)NH host
cells (Figs. 2 and 3B). Moreover, the BL21(DE3)NH host
phenotype was stable, since it continued to give rise to
small colonies on LB–carbenicillin agar plates upon
transformation with the recombinant plasmids tested.
Scanning electron microscopy (SEM) results show that
the E. coli BL21(DE3)NH cells have a distinct morphol-
ogy of long bacillary shape (F ig. 4A). The elongation is
evident and contrasts with the E. coli BL21(DE3) mor-
phology (Fig. 4B). The BL21(DE3)NH host cell is a
Gram-negative bacterium. Moreover, the biochemical
properties of BL21(DE3) NH host cells were determined
by standard methods for E. coli identification: urease
production, indole production, carbohydrate fermenta-
tion, b-galactos idase activity, citrate utilization, and
growth on MacConckey agar medium. All these tests
confirm that BL21(DE3)NH host cell is a strain of E. coli.
The mechanisms of translucent to opaque phenotype,
variations in colony morphology (large and small), and
alteration to the long bacillary shape of BL21(DE3)NH
cells remain unknown. It has been shown that opacity is
correlated with a robust transcription of the recombinant
gene whereas translation of recombinant RNA is not re-
quired [30]. Interestingly, uncoupling of transcription
from translation has been implicated in toxicity of re-
combinant protein expression leading to host cells pro-
ducing large amounts of the recombinant mRNA while
target proteins are maintained at rather low level s [16]. In
addition, expression of gene products that do not con-
tribute to the metabolism of bacterium, under the growth
conditions used in a particular experiment, has been
shown to result in cumulative breakdown of rRNAs and
ensuing loss of ribosomes and protein synthetic capacity
[31]. Although the mechanism that led the empirically
selected hos t to tolerate the recombinant mutant protein
expression is unknown, the BL21(DE3)NH cells derived
from small-opaque colonies transformed with re-
combinant plasmids resulted in small-opaque transfor-
mants, suggesting that the persistence of the phenotype
points to a stable genetic switch.
Leaky expression has been shown to occur in the pET
system [13,32,33]. The BL21(DE3)NH transformants
showed an absolute requirement for IPTG induction to
achieve expression of the recombinant proteins tested
(Figs. 2 and 3B). Accordingly, the genetic alteration of
the empirically selected host reported here appears to
have resulted in a tight control of the highly processive
T7 RNA polymerase that is under control of the IPTG-
inducible lacUV5 promoter. Moreover, the IPTG ab-
solute requirement for expression of the mycobacterial
ENR and KAR proteins in the BL21(DE3)NH host cell
suggests that no inherent defect in the transcriptional
activity of the T7 promoter is present.
It has been suggested that the mechanism responsible
for the differences in colony morphology is bacterial
growth rate [29]. Although we have also found this to be
generally true, any causal relationship between opacity
and growth rate seems to be unwarranted because it is
somewhat difficult to establish a quantitative relationship
between them based on a qualitative observation
(opacity).
Protein production and crystallization must be opti-
mized if structural genomics will ever reach its goal of
solving the three-dimensional structure of the whole
Fig. 3. SDS–PAGE (12%) of the soluble fraction of cell extracts of either BL21(DE3) or BL21(DE3)NH transformed with either pET-23a(+)
(control), pET-23a(+)::mabAK157A, or pET-23a(+)::mabAY153F recombinant plasmids. (A) Plasmids transformed into BL21(DE3). Lane 1,
control in the absence of IPTG; lane 2, control with 1 mM IPTG; lane M, protein molecular weight standards (same as Fig. 1); lane 3, K157A KAR
in the absence of IPTG; lane 4, K157A KAR with 1 mM IPTG; lane 5, Y153F KAR in the absence of IPTG; and lane 6, Y153F KAR with 1 mM
IPTG. (B) Plasmids transformed into BL21(DE3)NH. Lane 1, control in the absence of IPTG; lane 2, control with 1 mM IPTG; lane M, protein
molecular weight standards (same as Fig. 1); lane 3, K157A KAR in the absence of IPTG; lane 4, K157A KAR with 1 mM IPTG; lane 5, Y153F
KAR in the absence of IPTG; and lane 6, Y153F KAR with 1 mM IPTG. The arrow points to the band corresponding to the expected molecular
weight of M. tuberculosis KAR proteins (25.6 kDa).
S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125 123
proteome encoded by a given genome [34]. Unfortu-
nately, even when a genome can be sequenced, only up to
20% of the protein targets can produce soluble proteins
under very basic experimental conditions [35]. Thus, ex-
pression of proteins in soluble form has been identified as
an important bottleneck in efforts to determine biological
activity and crystal structure of M. tuberculosis proteins
[17]. A number of strategies for optimizing the yields of
heterologous protein expression in E. coli have been de-
scribed [36,37], including the use of different types of
promoters, placement of transcription terminators up-
stream of the promoter that drives expression of the gene
of interest, reduction of potential secondary-structure
formation at the 5
0
end of the transcript, use of transla-
tional enhancer, positioning of specific sequences in the 5
0
untranslated or 3
0
unstranslated regions of labile heter-
ologous mRNAs to prolong their half-lives, altering rare
codons in the target gene or coexpressing genes that en-
code rare tRNAs, mini mizing proteolysis by mutations to
avoid the E. coli ‘‘N-end rule’’ proteolytic pathway, and
altering fermentation conditions such as nutrient com-
position, temperature, and pH. Incidentally, although
wild-type KAR has nine rare codons (2Â CCC–proline,
5Â GGA–glycine, and 2Â ATA–isoleucine) and wild-
type ENR has six rare codons (2Â CCC–proline, 3Â
GGA–glycine, and 1Â AGG–arginine), they have no
adverse effect on the synthesis and yield of these re-
combinant proteins, which corroborates the conclusion
arrived by others [37] that the mere presence of rare co-
dons in a gene does not necessarily dictate poor transla-
tion of that gene.
The E. coli host reported here may represent an ad-
ditional tool to obtain mycobacterial soluble proteins
using T7-RNA polymerase-based systems as has been
the case for the C41(DE3) and C43(DE3) hosts
[15,16,25] now marketed by Avidis S.A. (France). Em-
pirical host selection requires no furt her genetic ma-
nipulation of recombinant plasmids and may yield
strains that overcome toxic effects associated with the
overexpression of heterologous protei ns. As poin ted out
by Miroux and Walker [16], even though the host se-
lection described here is empirical, it has the advantage
that it encompasses the entire complexity of the biology
of the expression system and provides a means of
modifying it. The expression of M. tuberculosis ENR
and KAR mutants will provide protein in quantities
necessary for determination of the mechanisms of action
of these enzymes by steady -state and pre-steady-state
kinetics as well as for crystallization trials aiming at X-
ray data collection. Enzymological and structural stud-
ies of M. tuberculosis ENR and KAR mutants should
help in the design of enzyme inhibitors to be tested as
new antitubercular agents.
Acknowledgments
Financial support for this work was provided by
Millennium Initiative Program MCT-CNPq, Ministry
of Health-Secretary of Health Policy (Brazi l) to D.S.S.
and L.A.B. D.S.S. and L.A.B. also acknowledge grants
awarded by PADCT, CNPq, an d FINEP. L.A.B.
(CNPq, 520182/99-5) is a researcher for the Brazilian
Council for Scientific and Technological Development.
We also thank Dr. Luiz Ant
^
onio Suita de Castro and
Embrapa Clima Temperado for the preparation of mi-
crographs and laboratory facilities.
References
[1] M.C. Raviglione, The TB epidemic from 1992 to 2002, Tubercu-
losis 83 (2003) 4–14.
Fig. 4. Scanning electron microscopy of E. coli hosts observed at
3000Â amplification. (A) E. coli BL21(DE3)NH and (B) E. coli
BL21(DE3).
124 S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125
[2] B.R. Bloom, C.J.L. Murray, Tuberculosis: commentary on a
reemergent killer, Science 257 (1992) 1055–1064.
[3] L.A. Basso, J.S. Blanchard, Resistance to antitubercular drugs,
Adv. Exp. Med. Biol. 456 (1998) 115–144.
[4] A. Plabos-M
endez, D.K. Gowda, T.R. Frieden, Controlling
multidrug-resistant tuberculosis and access to expensive drugs: a
rational framework, Bull. World Health Org. 80 (2002) 489–495.
[5] R.J. OÕBrien, P.P. Nunn, The need for new drugs against
tuberculosis, Am. J. Respir. Crit. Care Med. 162 (2001) 1055–
1958.
[6] S.T. Cole, R. Brosch, J. Parkhill, T. Garnier, C. Churcher, D.
Harris, S.V. Gordon, K. Eiglmeier, S. Gas, C.E. Barry III, F.
Tekaia, K. Badcock, D. Basham, D. Brown, T. Chillingworth, R.
Connor, R. Davies, K. Devlin, T. Feltwell, S. Gentles, N. Hamlin, S.
Holroyd, T. Hornsby, K. Jagels, A. Krogh, J. McLean, S. Moule, L.
Murphy, K. Oliver, J. Osborne, M.A. Quail, M.A. Rajandream, J.
Rogers, S. Rutter, K. Seeger, J. Skelton, S. Squares, R. Squares, J.E.
Sulston, K. Taylor, S. Whitehead, B.G. Barrell, B.G. Deciphering,
Deciphering the biology of Mycobacterium tuberculosis from the
complete genome sequence, Nature 393 (1998) 537–544.
[7] M.S. Glickman, W.R. Jacobs Jr., Microbial pathogenesis of
Mycobacterium tuberculosis: dawn of a discipline, Cell 104 (2001)
477–485.
[8] T.M. Shinnick, C.H. King, F.D. Quinn, Molecular biology,
virulence, and pathogenicity of mycobacteria, Am. J. Med. Sci.
309 (1995) 92–98.
[9] F. Baneyx, Recombinant protein expression in Escherichia coli,
Curr. Opin. Biotechnol. 10 (1999) 411–421.
[10] F.W. Studier, B.A. Moffat, Use of bacteriophage T7 RNA
polymerase to direct selective high-level expression of cloned
genes, J. Mol. Biol. 189 (1986) 113–130.
[11] A. Qu
emard, J.C. Sacchettini, A. Dessen, C. Vilch
eze, R. Bittman,
W.R. Jacobs Jr., J.S. Blanchard, Enzymatic characterization of
the target for isoniazid in Mycobacterium tuberculosis, Biochem-
istry 34 (1995) 8235–8241.
[12] L.A. Basso, D.S. Santos, W. Shi, R.H. Furneaux, P.C. Tyler, V.L.
Schramm, J.S. Blanchard, Purine nucleoside phosphorylase from
Mycobacterium tuberculosis analysis of inhibition by a transition-
state analogue and dissection by parts, Biochemistry 28 (2001)
8196–8203.
[13] J.S. Oliveira, C.A. Pinto, L.A. Basso, D.S. Santos, Cloning and
overexpression in soluble form of functional shikimate kinase and
5-enolpyruvylshikimate 3-phosphate synthase enzymes from My-
cobacterium tuberculosis, Protein Express. Purif. 3 (2001) 430–435.
[14] W. Shi, L.A. Basso, D.S. Santos, P.C. Tyler, R.H. Furneaux, J.S.
Blanchard, S.C. Almo, V.L. Schramm, Structures of purine
nucleoside phosphorylase from Mycobacterium tuberculosis in
complex with immucillin-H and its pieces, Biochemistry 28 (2001)
8204–8215.
[15] L. Kremer, K.M. Nampoothiri, S. Lesjean, L.G. Dover, S.
Graham, J. Betts, P.J. Brennan, D.E. Minnikin, C. Locht, G.S.
Besra, Biochemical characterization of acyl carrier protein
(AcpM) and malonyl-CoA:AcpM transacylase (mtFabD), two
major components of Mycobacterium tuberculosis fatty acid
synthase, J. Biol. Chem. 276 (2001) 27967–27974.
[16] B. Miroux, J.E. Walker, Over-production of proteins in Esche-
richia coli: mutant host that allow synthesis of some membrane
proteins and globular proteins at high levels, J. Mol. Biol. 260
(1996) 289–298.
[17] R. Vicentelli, C. Bignon, A. Gruez, S. Canaan, G. Sulzenbacher,
M. Tegoni, V. Campanacci, C. Cambillau, Medium-scale struc-
tural genomics: strategies for protein expression and crystalliza-
tion, Acc. Chem. Res. 36 (2003) 165–172.
[18] E.K. Schroeder, O.N. de Souza, D.S. Santos, J.S. Blanchard, L.A.
Basso, Drugs that inhibit mycolic acid biosynthesis in Mycobac-
terium tuberculosis, Curr. Pharm. Biotechnol. 3 (2002) 197–225.
[19] A. Banerjee, E. Dubnau, A. Qu
emard, V. Balasubramanian, K.S.
Um, T. Wilson, D. Collins, G. Lisle, W.R. Jacobs Jr., inhA, a gene
encoding a target for isoniazid and ethionamide in Mycobacterium
tuberculosis, Science 263 (1994) 227–230.
[20] A. Banerjee, M. Sugantino, J.C. Sacchettini, W.R. Jacobs Jr., The
mabA gene from the inhA operon of Mycobacterium tuberculosis
encodes a 3-ketoacyl reductase that fails to confer isoniazid
resistance, Microbiology 144 (1998) 2697–2704.
[21] P.J. Brennan, Structure, function, and biogenesis of the cell wall of
Mycobacterium tuberculosis, Tuberculosis 83 (2003) 91–97.
[22] D.A. Rozwarski, C. Vilch
eze, M. Sugantino, R. Bittman, J.C.
Sacchettini, Crystal structure of the Mycobacterium tuberculo-
sis enoyl-ACP reductase, InhA, in complex with NAD
þ
and a
C16 fatty acyl substrate, J. Biol. Chem. 274 (1999) 15582–
15589.
[23] S. Parikh, D.P. Moynihan, G. Xiao, P.J. Tonge, Roles of tyrosine
158 and lysine 165 in the catalytic mechanism of InhA, the enoyl-
ACP reductase from Mycobacterium tuberculosis, Biochemistry 38
(1999) 13623–13634.
[24] B. Persson, Y. Kallberg, U. Oppermann, H. J
ornvall, Coenzyme-
based functional assignments of short-chain dehydrogenases/
reductases (SDRs) family, Chem. Biol. Interact. 143-144 (2003)
271–278.
[25] J. Shanklin, Overexpression and purification of the Escherichia
coli inner membrane enzyme acyl acyl carrier protein synthase in
an active form, Protein Express. Purif. 18 (2000) 355–360.
[26] S.N. Ho, H.D. Hunt, R.M. Horton, J.K. Pullen, L.R. Pease, Site
directed mutagenesis by overlap extension using the polymerase
chain reaction, Gene 77 (1989) 51–59.
[27] M.A. Hayat, Basic Electron Microscopy Techniques, Nostrand
Press, New York, 1972.
[28] M.A. Hayat, Basic Techniques for Transmission Electron Mi-
croscopy, Academic Press, Orlando, 1986.
[29] R.C. Austin, D. Singh, P.C.Y. Liaw, H.J. Craig, Visual detection
method for identifying recombinant bacterial colonies, Bio.Tech-
niques 18 (1995) 380–384.
[30] S. Barik, Relationship between opacity of transformed E. coli
colonies and over-expression of the recombinant transcript,
BioTechniques 22 (1997) 112–118.
[31] H. Dong, L. Nilsson, C.G. Kurland, Gratuitous overexpression of
genes in Escherichia coli leads to growth inhibition and ribosome
destruction, J. Bacteriol. 177 (1995) 1497–1504.
[32] T.H. Grossman, E.S. Kawasaki, S.P. Punreddy, M.S. Osburne,
Spontaneous cAMP-dependent derepression of gene expression in
stationary phase plays a role in recombinant expression instability,
Gene 209 (1998) 95–103.
[33] K.C. Kelley, K.J. Huestis, D.A. Austen, C.T. Sanderson, M.A.
Donoghue, S.K. Stickel, E.S. Kawasaki, M.S. Osburne, Regula-
tion of sCD4-183 gene expression from phage-T7-based vectors in
Escherichia coli, Gene 156 (1995) 33–36.
[34] W.R.A. Taylor, A Ôperiodic tableÕ for protein structures, Nature
415 (2002) 657–660.
[35] S.A. Lesley, P. Kuhn, A. Godzik, A.M. Deacon, I. Mathews, A.
Kreusch, G. Spraggon, H.E. Klock, D. McMullan, T. Shin, J.
Vincent, A. Robb, L.S. Brinen, M.D. Miller, T.M. McPhillips,
M.A. Miller, D. Scheibe, J.M. Canaves, C. Guda, L. Jaroszewski,
T.L. Selby, M.A. Elsliger, J. Wooley, S.S. Taylor, K.O. Hodgson,
I.A. Wilson, P.G. Schultz, R.C. Stevens, Structural genomics of
the Thermotoga maritima proteome implemented in a high-
throughput structure determination pipeline, Proc. Natl. Acad.
Sci. USA 99 (2002) 11664–11669.
[36] G. Hannig, S.C. Makrides, Strategies for optimizing heterolo-
gous protein expression in E. coli, Trends Biotechnol. 16 (1998)
54–60.
[37] S.C. Makrides, Strategies for achieving high-level expression of
genes in Escherichia coli, Microbiol. Rev. 60 (1996) 512–538.
S.S. Poletto et al. / Protein Expression and Purification 34 (2004) 118–125 125
ANEXO III
57
Protein Expression and PuriWcation 40 (2005) 23–30
www.elsevier.com/locate/yprep
1046-5928/$ - see front matter 2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.pep.2004.06.040
DAHP synthase from Mycobacterium tuberculosis H37Rv: cloning,
expression, and puriW cation of functional enzyme
Caroline Rizzi
a
, Jeverson Frazzon
b
, Fernanda Ely
a
, Patrícia G. Weber
a
,
Isabel O. da Fonseca
a
, Michelle Gallas
a
, Jaim S. Oliveira
a
, Maria A. Mendes
c
,
Bibiana M. de Souza
c
, Mário S. Palma
c
, Diógenes S. Santos
d,¤
, Luiz A. Basso
a,¤
a
Departamento de Biologia Molecular e Biotecnologia, Universidade Federal do Rio Grande do Sul, Porto Alegre, RS 91501-970, Brazil
b
Departamento de Ciência dos Alimentos, Universidade Federal do Rio Grande do Sul, Porto Alegre, RS 91501-970, Brazil
c
Departamento de Biologia/CEIS, Universidade do Estado de São Paulo, Rio Claro, SP 13506-900, Brazil
d
Centro de Pesquisa e Desenvolvimento em Biologia Molecular e Funcional, PontíWcia Universidade Católica do Rio Grande do Sul,
Porto Alegre, RS 90619-900, Brazil
Received 13 April 2004, and in revised form 18 June 2004
Available online 8 December 2004
Abstract
Tuberculosis (TB), caused by Mycobacterium tuberculosis, remains the leading cause of mortality due to a bacterial pathogen.
According to the 2004 Global TB Control Report of the World Health Organization, there are 300,000 new cases per year of multi-
drug resistant strains (MDR-TB), deWned as resistant to isoniazid and rifampicin, and 79% of MDR-TB cases are now “super
strains,” resistant to at least three of the four main drugs used to treat TB. Thus there is a need for the development of eVective new
agents to treat TB. The shikimate pathway is an attractive target for the development of antimycobacterial agents because it has been
shown to be essential for the viability of M. tuberculosis, but absent from mammals. The M. tuberculosis aroG-encoded 3-deoxy-
D-
arabino-heptulosonate 7-phosphate synthase (mtDAHPS) catalyzes the Wrst committed step in this pathway. Here we describe the
PCR ampliWcation, cloning, and sequencing of aroG structural gene from M. tuberculosis H37Rv. The expression of recombinant
mtDAHPS protein in the soluble form was obtained in Escherichia coli Rosetta-gami (DE3) host cells without IPTG induction. An
approximately threefold puriWcation protocol yielded homogeneous enzyme with a speciWc activity value of 0.47 U mg
¡1
under the
experimental conditions used. Gel Wltration chromatography results demonstrate that recombinant mtDAHPS is a pentamer in solu-
tion. The availability of homogeneous mtDAHPS will allow structural and kinetics studies to be performed aiming at antitubercular
agents development.
2004 Elsevier Inc. All rights reserved.
Keywords: Mycobacterium tuberculosis; Shikimate pathway; DAHP synthase; Protein expression
Tuberculosis (TB)
1
remains the leading cause of mor-
tality due to a bacterial pathogen, Mycobacterium tuber-
culosis. The interruption of centuries of decline in case
rates of TB occurred, in most cases, in the late 1980s and
involved the USA and some European countries due to
increased poverty in urban settings and the immigration
*
Corresponding authors. Fax: +55 51 3166234.
E-mail addresses: diogenes@pucrs.br (D.S. Santos), labasso@dna.cbiot.ufrgs.br (L.A. Basso).
1
Abbreviations used: TB, tuberculosis; MDR-TB, multidrug-resistant; PEP, phosphoenolpyruvate; E4P, D-erythrose-4-phosphate; DAHPS, 3-de-
oxy-
D-arabino-heptulosonate 7-phosphate synthase; DMSO, dimethyl sulfoxide; LB, Luria–Bertani; SDS–PAGE, sodium dodecyl sulfate–polyacryl-
amide gel electrophoresis; IPTG, isopropyl -
D-thiogalactoside; -ME, -mercaptoethanol; DAHPS(Phe), phenylalanine-regulated DAHPS;
DAHPS(Try), tyrosine-regulated DAHPS; DAHPS(Trp), tryptophan-regulated DAHPS.
24 C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30
from TB high-burden countries [1]. Thus, no sustainable
control of TB epidemics can be reached in any country
without properly addressing the global epidemic. It is
estimated that 8.2 million new TB cases occurred world-
wide in the year 2000, with approximately 1.8 million
deaths in the same year, and more than 95% of those
were in developing countries [2]. Approximately, 2 bil-
lion individuals are believed to harbor latent TB based
on tuberculin skin test surveys [3], which represents a
considerable reservoir of bacilli. Possible factors under-
lying the resurgence of TB worldwide include the HIV
epidemic, increase in the homeless population, and
decline in health care structures and national surveil-
lance [4]. Another contributing factor is the evolution of
multi-drug resistant strains (MDR-TB), deWned as resis-
tant to isoniazid and rifampicin, which are the most
eVective Wrst-line drugs [5]. According to the 2004
Global TB Control Report of the World Health Organi-
zation, there are 300,000 new cases per year of MDR-TB
worldwide, and 79% of MDR-TB cases are now “super
strains,” resistant to at least three of the four main drugs
used to treat TB [6]. The factors that most inXuence the
emergence of drug-resistant strains include inappropri-
ate treatment regimens, and patient noncompliance in
completing the prescribed courses of therapy due to the
lengthy standard “short-course” treatment or when the
side eVects become unbearable [7]. Hence, faster acting
and eVective new drugs to better combat TB, including
MDR-TB, are needed.
The shikimate pathway is an attractive target for the
development of herbicides and antimicrobial agents
because it is essential in algae, higher plants, bacteria,
and fungi, but absent from mammals [8]. In mycobacte-
ria, the shikimate pathway leads to the biosynthesis of
chorismic acid, which is a precursor for the synthesis of
aromatic amino acids, naphthoquinones, menaquinones,
and mycobactins [9]. The salicylate-derived mycobactin
siderophores have been shown to be essential for M.
tuberculosis growth in macrophages [10]. More recently,
the shikimate pathway has been shown by disruption of
aroK gene, which codes for the shikimate kinase enzyme,
to be essential for the viability of M. tuberculosis [11].
The absence from the human host and essentiality of
mycobacterial shikimate pathway indicate that any of its
enzymes are promising targets for the development of
potentially non-toxic antimycobacterial agents.
Homologues to enzymes in the shikimate pathway
have been identiWed in the genome sequence of M. tuber-
culosis [12]. The Wrst committed step in the shikimate
pathway is catalyzed by 3-deoxy-
D-arabino-heptuloson-
ate 7-phosphate (DAHP) synthase (DAHPS; EC
4.1.2.15). DAHPS catalyzes the stereospeciWc condensa-
tion of phosphoenolpyruvate (PEP) and
D-erythrose
4-phosphate (E4P), forming DAHP and inorganic phos-
phate [13]. Based on phylogenetic analysis, DAHPS has
been divided into two classes, class I and class II [14].
Escherichia coli expresses three DAHPS isoenzymes that
are representative of class II and require divalent metal
for activity [15], which play a role in catalysis and/or
structural integrity [16]. Each isoenzyme is speciWcally
inhibited by one of the three aromatic amino acids [8].
DAHPS(Phe), a homotetramer encoded by the aroG
gene, is feedback inhibited by phenylalanine; the aroH-
encoded DAHPS(Trp) and aroF-encoded DAHPS(Tyr)
are homodimers feedback inhibited by, respectively,
tryptophan and tyrosine. In M. tuberculosis genome,
however, only the aroG (Rv2178c) encoded DAHPS iso-
enzyme (mtDAHPS) has been proposed to be present by
sequence homology.
To determine the mechanism of action of mtDAHPS
by steady-state and pre-steady-state kinetics as well as
for X-ray crystal structure determination aiming at the
rational design of antimycobacterial agents, expression
of aroG encoded mtDAHPS in functional form and in
large quantity are needed. Accordingly, we here
describe the PCR ampliWcation, cloning, sequencing,
expression in E. coli Rosetta-gami (DE3) cells, puriWca-
tion to homogeneity, oligomeric state determination,
and assay of mtDAHP enzyme activity. Measurements
of enzyme activity conWrm the correct assignment to
the structural gene encoding mtDAHPS in M. tubercu-
losis. The availability of mtDAHPS will allow enzyme
kinetics and structural studies to be undertaken to pro-
vide a framework on which to base the design of new
agents with antitubercular activity with, hopefully, low
toxicity.
Materials and methods
PCR ampliWcation and cloning of M. tuberculosis aroG
gene
The design of synthetic oligonucleotide primers used
for PCR ampliWcation of aroG gene (5Ј-ggacatatgaactgg
accgtcgacatac-3Ј and 5Ј-cggatcctcagtcccgcagcatctccgc-3Ј)
was based on the complete genome sequence of M.
tuberculosis H37Rv [12]. These primers were comple-
mentary to, respectively, the amino-terminal coding and
carboxy-terminal noncoding strands of aroG gene con-
taining 5ЈNdeI and 3ЈBamHI restrictions sites, which are
in bold. This pair of primers was used to amplify the M.
tuberculosis aroG gene (1389 bp) from genomic DNA
using standard PCR conditions and the enzyme Pfu
DNA polymerase (Stratagene), which is a thermostable
polymerase that exhibits low error rate, thus lowering
the likelihood of introducing unwanted mutations. PCR
ampliWcation required the presence of 10% of dimethyl
sulfoxide (DMSO) in the reaction mixture. The PCR
product was puriWed by electrophoresis on low melting
agarose, digested with NdeI and BamHI (Boehringer–
Mannheim), and cloned into pET-23a(+) (Novagen)
C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30 25
expression vector, which had previously been digested
with the same restriction enzymes. To both conWrm the
identity of the cloned gene and ensure that no mutations
were introduced by the PCR ampliWcation step, the
DNA sequence of the ampliWed M. tuberculosis aroG
structural gene was determined by dideoxy-chain termi-
nation method [17], using the Thermo Sequenase radio-
labeled terminator cycle sequencing kit (Amersham
Biosciences).
Expression of mtDAHPS
The recombinant plasmid pET-23a(+)::aroG was
transformed into electrocompetent E. coli Rosetta-gami
(DE3) cells (Novagen), and selected on LB agar plates
containing 50 gmL
¡1
carbenicillin, 15 gmL
¡1
kana-
mycin, 34 gmL
¡1
chloramphenicol, and 12.5 gmL
¡1
tetracycline. Single colonies were used to inoculate
500 mL LB medium, containing the same antibiotics and
concentrations of LB solid medium, and grown at 37 °C
and 180 rpm for 24 h, without addition of isopropyl -
D-
thiogalactopyranoside (IPTG). Cells were harvested by
centrifugation at 48,000g for 20 min at 4 °C, and stored
at ¡20 °C. For protein expression analysis, 10 mg of
stored cells was resuspended in 500 L BuVer A (50 mM
Tris–HCl, pH 7.8), disrupted by soniWcation, and cell
debris was removed by centrifugation. Both soluble and
insoluble fractions were analyzed by SDS–PAGE 12%
[18]. Control experiments were performed under the
same experimental conditions except that E. coli host
cells were transformed with the expression vector lack-
ing the target gene.
PuriWcation of recombinant mtDAHPS
Approximately, 36 g of cells was collected by centri-
fugation (48,000g for 20 min) from 6 L of LB medium.
All subsequent steps were performed on ice or at 4 °C.
Frozen cells (36 g) were thawed and resuspended in
BuVer A (4 mL of buVer per gram of cell paste) contain-
ing 1 mM -mercaptoethanol (-ME) (Sigma) and
0.2 mg mL
¡1
of lysozyme, and the mixture was stirred
for 30 min. Cells were disrupted by sonication, and cell
debris was removed by centrifugation (48,000g for
30 min). The supernatant was incubated with 1% w/v of
streptomycin sulfate for 15 min, and centrifuged
(48,000g for 30 min). Solid ammonium sulfate was
added to the supernatant fraction to a concentration of
25% saturation, incubated for 30 min, and centrifuged
as above. The resultant pellet was resuspended in 70 mL
BuVer A containing 1 mM -ME and dialyzed against
three changes of 2 L of the same buVer using a dialysis
tubing with molecular weight cut-oV of 12,000–4000 Da.
The sample was clariWed by centrifugation and loaded
on a Q-Sepharose Fast Flow (2.6 cm £ 8.2 cm) anion
exchange column (Amersham Biosciences) previously
equilibrated with BuVer A and fractionated using a
600 mL 0.0–0.6 M NaCl linear gradient. The fractions
containing mtDAHPS (0.35–0.38 M NaCl) were pooled,
concentrated to 8.0 mL using an Amicon ultraWltration
cell (MW 30,000 Da), and loaded on a Sephacryl S-200
HR (2.6 cm £ 60 cm) gel Wltration column (Amersham
Biosciences) at 0.5 mL min
¡1
. The protein was eluted
with BuVer B (BuVer A containing 200 mM NaCl) at the
same Xowrate. The active fractions were loaded on a
Mono Q HR 16/10 anion exchange column (Amersham
Biosciences) equilibrated with BuVer A and eluted with
400 mL linear 0.0–0.6 M NaCl gradient. The active frac-
tions, which exhibited a single band on SDS–PAGE,
were pooled, quickly frozen in liquid nitrogen, and
stored at ¡80 °C.
Determination of protein concentration
Protein concentrations were determined using the
Bio-Rad Laboratories protein assay kit (Bradford
method) [19] and bovine serum albumin as standard.
Determination of mtDAHPS molecular mass
The molecular mass of native mtDAHPS homoge-
neous protein was determined by gel Wltration chroma-
tography using a Sephacryl S-200 (HR 10/30)
(Amersham Biosciences) equilibrated with BuVer B at a
Xowrate of 0.4 mL min
¡1
. Protein molecular mass stan-
dards were from Gel Filtration LMW and HMW Cali-
bration Kit from Amersham Biosciences. Protein elution
was monitored at 280 nm.
mtDAHPS enzyme assay
Enzyme activity of recombinant mtDAPHS protein
was assayed in the forward direction by a continuous
spectrophotometric method described by Shoner and
Hermann [20], monitoring the decrease in phosphoenol-
pyruvate (PEP) concentration at 232 nm
( D 2.8 £ 10
3
M
¡1
cm
¡1
) on a Multi-Spec 1501 photodi-
ode array spectrophotometer (Shimadzu). All reactions
were carried out at 25 °C and initiated with addition of
enzyme to a reaction mixture containing: 50mM Tris–
HCl, pH 7.0, 400 M E4P (Sigma), 1 mM -ME, and
200 M PEP (Acrós Organics) in a total volume of
500 L. One unit of enzyme activity (U) is deWned as the
amount of enzyme catalyzing the conversion of 1 mol
PEP/min at 25 °C.
N-terminal amino acid sequencing
The N-terminal amino acid residues of homogeneous
recombinant mtDAHPS were identiWed by automated
Edman degradation sequencing using a PPSQ 21A gas-
phase sequencer (Shimadzu).
26 C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30
Mass spectrometry analysis
The homogeneity of recombinant protein preparation
was assessed by mass spectrometry (MS), employing
some adaptations made to the system described by
Chassaigne and Lobinski [21]. Samples were analyzed on
a triple quadrupole mass spectrometer, model QUAT-
TRO II, equipped with standard electrospray (ESI)
probe (Micromass, Altrinchan), adjusted to ca.
250 Lmin
¡1
. The source temperature (80 °C) and needle
voltage (3.6kV) were maintained constant throughout
the experimental data collection, applying a drying gas
Xow (nitrogen) of 200 Lh
¡1
and a nebulizer gas Xow of
20 Lh
¡1
. The mass spectrometer was calibrated with
intact horse heart myoglobin and its typical cone-volt-
age induced fragments. The subunit molecular mass of
recombinant protein mtDAHPS was determined by ESI-
MS, adjusting the mass spectrometer to give a peak with
at half-height of 1 mass unit, and the cone sample to
skimmer lens voltage controlling the ion transfer to mass
analyzer was set to 38 V. About 50 pmol (10 L) of each
sample was injected into the electrospray transport sol-
vent. The ESI spectrum was obtained in the multi-chan-
nel acquisition mode, scanning from 500 to 1800 m/z at a
scan time of 7 s. The mass spectrometer is equipped with
MassLynx and Transform software for data acquisition
and spectra handling.
Results and discussion
The PCR ampliWcation of aroG structural gene from
M. tuberculosis H37Rv genomic DNA required the pres-
ence of 10% DMSO in the reaction mixture (data not
shown). DMSO is a cosolvent that improves GC-rich
DNA denaturation and helps to overcome the diYculties
of polymerase extension through secondary structures,
altering the structural conformation of DNA templates
[22]. This result is consistent with the 65.6% G + C con-
tent of M. tuberculosis H37Rv genome [12]. PCR frag-
ment was inserted into pET23a(+) expression vector [23]
between NdeI and BamHI restriction sites. DNA
sequencing of the entire aroG structural gene by the dide-
oxy chain termination method both conWrmed the iden-
tify of the cloned PCR product and showed that no
mutations were introduced by the DNA ampliWcation
step.
Recombinant plasmids were introduced into E. coli
BL21(DE3) host cells by electroporation. Unfortunately,
recombinant mtDAHPS remained in the insoluble frac-
tion (Fig. 1). Since one of the goals of the present work
was to conWrm the correct assignment to the structural
gene encoding mtDAHPS, eVorts were made to express
recombinant M. tuberculosis DAHPS in its soluble,
active form avoiding unfolding and refolding protocols
because they cannot guarantee that they will yield large
amounts of biologically active product [24]. In addition,
a number of protocols were tested to obtain mtDAHPS
in the soluble fraction to no avail, including buVer addi-
tives (urea, deoxycholic acid, Triton X-100, and high
NaCl concentrations) and reduced cultivation tempera-
ture (20, 25, and 30 °C). In practice, it is usually worth-
while to test several diVerent vector/host combinations
to obtain the best possible yield of protein in its desired
form. Accordingly, a number of commercially available
strains of E. coli host cells were tested in an attempt to
produce mtDAHPS in the soluble fraction. Analysis of
the relationship between codon preference and expres-
sion level led to the classiWcation of E. coli genes into
three main classes [25]. Class II genes, which correspond
to genes highly and continuously expressed during expo-
nential growth that is likely to resemble the tRNA popu-
lation available for recombinant protein expression,
have a number of avoided codons with frequencies of
less than 6%. InsuYcient tRNA pools can lead to prema-
ture translational termination, translation frameshifting
or amino acid misincorporation that might result in
expression of nonproperly folded recombinant protein
[26]. Rare codons near the N-terminus of a coding
sequence can have a severe eVect on heterologous
expression in E. coli [27]. Four rare codons for heterolo-
gous gene expression in E. coli are present near the N-
terminus of M. tuberculosis aroG structural gene
(1 £ AUA for isoleucine, 3 £ CCC for proline). To test
whether these rare codons may have any eVect on aroG
expression, E. coli Rosetta (DE3) strain harboring
tRNA genes for AGG, AGA, AUA, CUA, CCC, and
Fig. 1. SDS–PAGE (12%) of the soluble and insoluble fractions of the
cell extracts of either BL21(DE3) or Rosetta-gami (DE3) host cells
transformed with either pET-23a(+) (control) or pET-23a(+)::aroG.
Expression conditions were 24 h at 37 °C without IPTG addition. Lane
1: insoluble fraction of BL21 (DE3) transformed with pET-23(+); lane
2: insoluble fraction of BL21(DE3) transformed with pET-
23(+)::aroG; lane 3: soluble fraction of BL21(DE3) transformed with
pET-23(+); lane 4: soluble fraction of BL21(DE3) transformed with
pET-23(+)::aroG; lane 5: MW marker “High range” (Gibco-BRL);
lane 6: insoluble fraction of Rosetta-gami (DE3) transformed with
pET-23(+); lane 7: insoluble fraction of Rosetta-gami (DE3) trans-
formed with pET-23(+)::aroG; lane 8: soluble fraction of Rosetta-gami
(DE3) transformed with pET-23(+); and lane 9: soluble fraction of
Rosetta-gami (DE3) transformed with pET-23(+)::aroG. Molecular
mass of mtDAHPS is approximately 50.6 kDa.
C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30 27
GGA rare codons on a chloramphenicol-resistant plas-
mid [28] was transformed with pET-23a(+)::aroG recom-
binant plasmid. Disappointingly, recombinant
mtDAHPS remained in the insoluble fraction (data not
shown), thereby discarding any eVect of the mycobacte-
rial aroG rare codons on recombinant protein expres-
sion. Although E. coli DAHPS has no disulWde bridges
despite possessing three cysteine residues, there is no
experimental evidence for the absence of disulWde
bridges in mtDAHPS, which possesses Wve cysteine resi-
dues (Cys 87, Cys 231, Cys 365, Cys 420, and Cys 440),
and a less reducing cytoplasmatic environment could
improve mtDAHPS solubility. The Origami E. coli host
strains (Novagen) have mutations in both the thiore-
doxin reductase (trxB) and glutathione reductase (gor)
genes, which greatly enhances disulWde bond formation
in the cytoplasm [29,30]. Unfortunately, none of the pro-
tocols tested yielded soluble mtDAHPS. The Rosetta-
gami (DE3) E. coli host strain (Novagen) combines the
features of Rosetta and Origami strains. SDS–PAGE
analysis showed that expression of recombinant
mtDAHPS protein in its soluble form with the expected
molecular mass (»51 kDa) could be achieved using the
Rosetta-gami (DE3) cells grown at 37 °C for 24 h with
no IPTG induction (Fig. 1). The underlying reason for
this result is unclear; however, it underscores the need
for optimization of vector/host combinations to achieve
soluble recombinant protein expression before attempt-
ing any unfolding/refolding protocols.
It should be pointed out that a screening of experi-
mental conditions was carried out to obtain high yield of
recombinant protein expression, including temperature
of growth, culture aeration, medium type, hours of
growth after IPTG induction, and hours of growth in the
absence of IPTG. The best results were obtained from
Rosetta-gami (DE3) E. coli cells grown for 24 h at 37 °C
in LB medium without IPTG induction as described
above. In the pET vector system (Novagen), target genes
are positioned downstream of the bacteriophage T7 late
promoter. Typically, production hosts contain a pro-
phage (DE3) encoding the highly processive T7 RNA
polymerase under control of the IPTG-inducible lacUV5
promoter that would ensure tight control of recombi-
nant gene basal expression [31,32]. In agreement with the
results presented here, leaky expression has been shown
to occur in the pET system [33–37]. It has been proposed
that leaky protein expression is a property of the lac-
controlled system as cells approach stationary phase in
complex medium and that cyclic AMP, acetate, and low
pH are required to achieve high-level expression in the
absence of IPTG induction, which may be part of a gen-
eral cellular response to nutrient limitation [38].
Enzyme activity measurements demonstrated that
there was a 92-fold increase in speciWc activity for
mtDAHPS when Rosetta-gami (DE3) E. coli harboring
either pET-23a(+)::aroG or pET-23a(+) crude extracts
were compared (Table 1), indicating that mtDAHPS was
expressed in its soluble and functional form. The puriW-
cation protocol of recombinant mtDAHPS, protein
determination, enzyme assay, and SDS–PAGE analysis
were as described in Materials and methods. Recombi-
nant mtDAHPS enzyme was puriWed approximately 3-
fold (Table 2) to electrophoretic homogeneity (Fig. 2).
The puriWcation protocol yielded approximately 5 mg of
homogeneous protein. A signiWcant loss in protein yield
occurred in the 25% ammonium sulfate precipitation
step (Table 2) because some mtDAHPS remained in the
supernatant. Protein precipitations by higher ammo-
nium sulfate concentrations were also carried out, yield-
ing larger amounts of recombinant mtDAHPS in the
pellet (data not shown). However, a number of contami-
nants co-precipitated with mtDAHPS. In particular, a
contaminant that co-eluted in subsequent chromato-
graphic steps when larger than 25% ammonium sulfate
concentrations were used. Accordingly, the 25% ammo-
nium sulfate precipitation step was deemed more appro-
priate for the puriWcation protocol because a signiWcant
amount of contaminants remained in the supernatant,
while mtDAHPS with a lower protein-contaminating
background remained in the pellet thus making the
Table 1
Measurements of recombinant DAHPS enzyme activity
a
Crude cell extract in 50 mM Tris–HCl, pH 7.0.
b
UmL
¡1
/mg mL
¡1
.
Cell extract
a
SpeciWc activity
b
(SA, U mg
¡1
) AS cloned/SA control
Control 0.0018 1
DAHPS 0.1651 92
Table 2
PuriWcation of M. tuberculosis 3-deoxy-
D-arabino-heptulosonate 7-phosphate synthase expressed in E. coli Rosetta-gami(DE3) transformed with
pET-23a(+)::aroG
a
a
Typical puriWcation protocol starting from 36 g wet weight cells obtained from 6 L of culture.
b
UmL
¡1
/mg mL
¡1
.
PuriWcation step Total protein (mg) Total enzyme activity (U) SpeciWc activity
b
(U mg
¡1
)PuriWcation fold Yield (%)
Crude extract 2442.38 403.23 0.17 1.00 100
Ammonium sulfate 136.78 44.11 0.32 1.95 11
Q-Sepharose Fast Flow 21.42 7.00 0.33 1.98 2
Sephacryl S-200 HR 17.83 4.48 0.25 1.52 1
Mono Q HR 16/10 4.77 2.24 0.47 2.84 0.6
28 C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30
subsequent puriWcation steps less demanding. The
samples of each puriWcation step were assayed for
DAHPS enzyme activity and compared to control exper-
iments to demonstrate that the observed values are
actual velocities of mtDAHPS activity. Enzyme activity
of homogeneous mtDAHPS was linearly dependent on
sample volume added to the reaction mixture (Fig. 3),
thereby showing that the initial velocity is proportional
to total enzyme concentration and that true initial veloc-
ities are being measured. E. coli DAHPS(Phe) has been
shown to be sensitive to oxidation, leading to inactiva-
tion [39]. Moreover, higher mtDAHPS enzyme activity
values could be observed in the presence of -ME (data
not shown). Accordingly, -ME was used in all steps of
the mtDAHPS puriWcation protocol. The mtDAHPS
speciWc activity has been found to be stable for at least
two months when stored at ¡80 °C.
The subunit molecular mass of active mtDAHPS was
determined to be 50510.38 Da by electrospray ionization
mass spectrometry (ESI-MS), consistent with the post-
translational removal of the N-terminal methionine resi-
due from the full length gene product (predicted mass:
50641.51 Da). The ESI-MS result also revealed no peak
at the expected mass for the three isoforms of E. coli
DAHPS enzymes (38804.03, 38735.18, and 38009.53 Da),
thus providing evidence for both the identity and purity
of the recombinant protein. The Wrst 11 N-terminal
amino acid residues of the recombinant protein were
identiWed as NWTVDIPIDQL by the Edman degrada-
tion chemistry protocol. This result unambiguously iden-
tiWes the homogeneous recombinant protein as
mtDAHPS and conWrms removal of the N-terminal
methionine residue from it. A common type of co-/post-
translational modiWcation of proteins synthesized in
prokaryotic cells is modiWcation at their N-termini.
Methionine aminopeptidase catalyzed cleavage of initia-
tor methionine is usually directed by the penultimate
amino acid residues with the smallest side chain radii of
gyration (glycine, alanine, serine, threonine, proline,
valine, and cysteine) [40]. The N-terminal methionine was
removed from the E. coli expressed recombinant
mtDAHPS enzyme, consistent with the Wnding that some
middle-sized penultimate amino acid residues (Asn, Asp,
Leu, and Ile) undergo N-terminal processing [41].
A value of 253 § 25 kDa was determined for the
molecular mass of native mtDAHPS homogeneous pro-
tein by analytical gel Wltration chromatography (data
not shown), suggesting that mtDAHPS is a pentamer in
solution. Whereas E. coli DAHPS(Phe) is a tetramer [42]
and E. coli DAHPS(Trp) is a dimer [43]. Interestingly,
more recently, recombinant DAHPS from Pyrococcus
furiosus has been shown to be a dimer in solution and
not to be inhibited by phenylalanine, tyrosine, or trypto-
phan [44].
DAHPS in most, not all, microorganisms is the target
for pathway regulation by negative feedback inhibition,
which controls carbon Xow into the shikimate pathway.
The most intensively investigated microorganism
DAHPS has been the E. coli enzyme, which possesses
three isoenzymes, each speciWcally regulated by one of
three aromatic amino acid end products, either Phe, Tyr,
or Trp [45]. The three isoforms have a common require-
ment for a metal cofactor, which can be similarly satis-
Wed by a range of divalent metal ions [46]. DAHPS
enzymes from a number of microorganisms have been
studied, such as Corynebacterium glutamicum [47], Ther-
motoga maritima [48], Bacillus subtilis [49], Saccharomy-
ces cerevisiae [50], and P. furiosus [44]. However, to the
best of our knowledge, this is the Wrst report on cloning,
expression, and puriWcation of functional DAHPS from
M. tuberculosis.
Fig. 2. SDS–PAGE analysis of pooled fractions from the puriWcation
steps of mtDAHPS. Lane 1, MW marker “High range” (Gibco-BRL);
lane 2, crude extract; lane 3, ammonium sulfate precipitation; lane 4,
Q-Sepharose Fast Flow ion exchange; lane 5, S-200 gel Wltration; and
lane 6, Mono Q ion exchange.
Fig. 3. Linear dependence of mtDAHPS activity on homogeneous pro-
tein volume. The rates of enzyme activity were performed in the for-
ward direction by continuously monitoring the decrease of
phosphoenolpyruvic acid at 232 nm.
C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30 29
Homogeneous mtDAHPS protein will provide pro-
tein in quantities necessary for studies on the enzyme
mechanism of action by steady-state and pre-steady-
state kinetics, its metal requirement, if any, and feedback
inhibition by Phe, Tyr, and Trp. The expression and
puriWcation of mtDAHPS reported here will also provide
protein for crystallization trials aiming at three-dimen-
sional structure determination by X-ray diVraction. The
three-dimensional structures of four forms of the Phe-
regulated isoenzyme of E. coli DAHPS have been solved
by X-ray crystallography [42,51–53], which should facili-
tate screening of experimental conditions to obtain crys-
tals of mtDAHPS in complex with its substrates,
possible metal cofactor and feedback inhibitors, if any.
Expression of functional proteins in soluble form has
been identiWed as an important bottleneck in eVorts to
determine biological activity and crystal structure of M.
tuberculosis proteins [54]. We hope that the results
reported here will contribute to eVorts towards the struc-
ture determination of potential targets in M. tuberculo-
sis. The enzymological and structural studies on
mtDAHPS should help in the design of enzyme inhibi-
tors to be tested as antimycobacterial agents.
Acknowledgments
Financial support for this work was provided by Mil-
lennium Initiative Program MCT-CNPq, Ministry of
Health-Department of Science and Technology-UNE-
SCO (Brazil) to D.S.S. and L.A.B. D.S.S. and L.A.B. also
acknowledge grants awarded by PADCT, CNPq, and
FINEP. L.A.B. (CNPq, 520182/99-5), D.S.S. (CNPq,
304051/1975-06), and M.S.P. (CNPq, 300337/2003-50)
are researchers awardees from the National Council for
ScientiWc and Technological Development of Brazil.
M.A.M. is a post-doctoral fellow from FAPESP (01/
05060-4).
References
[1] M.C. Raviglione, The TB epidemic from 1992 to 2002, Tuberculo-
sis 83 (2003) 4–14.
[2] E.L Corbett., C.J. Watt, N. Walker, D. Maher, B.G. Williams, M.C.
Raviglione, C. Dye, The growing burden of tuberculosis: global
trends and interactions with the HIV epidemic, Arch. Intern. Med.
1639 (2003) 1009–1021.
[3] C. Dye, S. Scheele, P. Dolin, V. Pathania, M.C. Raviglione, Global
burden of tuberculosis: estimated incidence, prevalence, and mor-
tality by country, JAMA 282 (1999) 677–686.
[4] B.R. Bloom, C.J.L. Murray, Tuberculosis: commentary on a
reemergent killer, Science 257 (1998) 1055–1064.
[5] L.A. Basso, J.S. Blanchard, Resistance to antitubercular drugs,
Adv. Exp. Med. Biol. 456 (1998) 115–144.
[6] World Health Organization, Anti-tuberculosis drug resistance in
the world: Third Global Report, 2004.
[7] K. Duncan, Progress in TB drug development and what is still
needed, Tuberculosis 83 (2003) 201–207.
[8] R. Bentley, The shikimate pathway—a metabolic tree with many
branches, Crit. Rev. Biochem. Mol. Biol. 25 (1990) 307–384.
[9] C. Ratledge, Nutrition, growth and metabolism, in: C. Ratledge,
J.L. Stanford (Eds.), The Biology of the Mycobacteria, vol. 1, Aca-
demic Press, London, 1982, pp. 185–271.
[10] J.J. de Voos, K. Rutter, B.G. Schroder, H. Su, Y. Zhu, C.E. Barry
III, The salicylate-derived mycobactin siderophores of Mycobac-
terium tuberculosis are essential for growth in macrophages, Proc.
Natl. Acad. Sci. USA 97 (2000) 1252–1257.
[11] T. Parish, N.G. Stoker, The common aromatic amino acid biosyn-
thesis pathway is essential in Mycobacterium tuberculosis, Micro-
biology 148 (2002) 3069–3077.
[12] S.T. Cole, R. Brosch, J. Parkhill, T. Garnier, C. Churcher, D. Har-
ris, S.V. Gordon, K. Eiglmeier, S. Gas, C.E. Barry III, F. Tekaia, K.
Badcock, D. Basham, D. Brown, T. Chillingworth, R. Connor, R.
Davies, K. Devlin, T. Feltwell, S. Gentles, N. Hamlin, S. Holroyd,
T. Hornsby, K. Jagels, B.G. Barrell, Deciphering the biology of
Mycobacterium tuberculosis from the complete genome sequence,
Nature 393 (1998) 537–544.
[13] P.R. Srinivasan, D.B. Sprinson, 2-Keto-3-deoxy-
D-arabo-heptonic
acid 7-phosphate synthase, J. Biol. Chem. 234 (1959) 716–722.
[14] M.R. Birck, R.W. Woodard, Aquifex aeolicus 3-deoxy-
D-manno-2-
octulosonic acid 8-phosphate synthase: a new class of KDO 8-P
synthase?, J. Mol. Evol. 52 (2001) 205–214.
[15] C.M. Stephens, R. Bauerle, Analysis of the metal requirement of 3-
deoxy-
D-arabino-heptulosonate-7-phosphate synthase from Esch-
erichia coli, J. Biol. Chem. 266 (1991) 20810–20817.
[16] P.A. Jordan, S. Bohle, C. Ramilo, J.N.S. Evans, New insights into
the metal center of 3-deoxy-
D-arabino-heptulosonate 7-phosphate
synthase, Biochemistry 40 (2001) 8387–8396.
[17] F. Sanger, S. Nicklen, R. Coulson, DNA sequencing with chain-ter-
minating inhibitors, Pro. Natl. Acad. Sci. USA 74 (1977) 5436–5467.
[18] U.K. Laemmli, Cleavage of structural proteins during the assem-
bly of the head of bacteriophage T4, Nature 227 (1970) 680–685.
[19] M.M. Bradford, R.A. McRorie, W.L. Williams, A rapid and sensi-
tive method for the quantiWcation of microgram quantities of pro-
tein utilizing the principle of protein-dye binding, Anal. Biochem.
72 (1976) 248–254.
[20] R. Shoner, K.M. Hermann, 3-Deoxy-
D-arabino-heptulosonate 7-
phosphate synthase, J. Biol. Chem. 252 (1976) 5440–5447.
[21] H. Chassaigne, R. Lobinski, Characterization of horse kidney
metallothionein isoforms by electrospray MS and reversed-phase
HPLC-electrospray MS, Analyst 123 (1998) 2125–2130.
[22] D. Pomp, J.F. Medrano, Organic solvents as facilitators of poly-
merase chain reaction, Biotechniques 10 (1991) 58–59.
[23] F.W. Studier, A.H. Rosenberg, J.J. Dunn, J.W. DubendorV, Use of
T7 RNA polymerase to direct expression of cloned genes, Meth-
ods Enzymol. 185 (1990) 60–89.
[24] C.H. Schein, Production of soluble recombinant proteins in bacte-
ria, Bio/technology 7 (1989) 1141–1149.
[25] A. Hénaut, A. Danchin, Analysis and predictions from Escherichia
coli sequences, or E. coli in silico, in: F.C. Neidhardt (Ed.), Esche-
chichia coli and Salmonella: Cellular and Molecular Biology, ASM
Press, Washington, DC, 1996, pp. 2047–2066.
[26] C. Kurland, J. Gallant, Error of heterologous protein expression,
Curr. Opin. Biotechnol. 7 (1996) 489–493.
[27] Y. Nakamura, T. Gojobori, T. Ikemura, Codon usage tabulated
from international DNA sequence databases: status for the year
2000, Nucleic Acids Res. 28 (2000) 292.
[28] R. Novy, D. Drott, K. Yaeger, R. Mierendorf, Overcoming the
codons bias of Escherichia coli for enhanced protein expression,
Innovations 12 (2001) 1–3.
[29] W.A. Prinz, F. Aslund, A. Holmgren, J. Beckwith, The role of the
thioredoxin and glutaredoxin pathways in redox-facilitate
attempts at crystallizing and elucidating the protein disulWde
bonds in the Escherichia coli cytoplasm, J. Biol. Chem. 272 (1997)
15661–15667.
30 C. Rizzi et al. / Protein Expression and PuriWcation 40 (2005) 23–30
[30] F. Baneyx, Recombinant protein expression in Escherichia coli,
Curr. Opin. Biotechnol. 10 (1999) 411–421.
[31] F.W. Studier, B.A. MoVatt, Use of bacteriophage T7 RNA poly-
merase to direct selective high-level expression of cloned genes, J.
Mol. Biol. 189 (1986) 113–130.
[32] J.W. DubendorV, F.W. Studier, Controlling basal expression in an
inducible T7 expression system by blocking the target T7 pro-
moter with lac repressor, J. Mol. Biol. 219 (1991) 45–59.
[33] L.A. Basso, D.S. Santos, W. Shi, R.H. Furneaux, P.C. Tyler, V.L.
Schramm, J.S. Blanchard, Purine nucleoside phosphorylase from
Mycobacterium tuberculosis. Analysis of inhibition by a transition-
state analogue and dissection by parts, Biochemistry 40 (2001)
8196–8203.
[34] J.S. Oliveira, C.A. Pinto, L.A. Basso, D.S. Santos, Cloning and
overexpression in soluble form of functional shikimate kinase and
5-enoylpyruvylshikimate 3-phosphate synthase enzymes from
Mycobacterium tuberculosis, Protein Expr. Purif. 3 (2001) 430–435.
[35] M.L.B. Magalhães, C.P. Pereira, L.A. Basso, D.S. Santos, Cloning
and expression of functional shikimate dehydrogenase (EC
1.1.1.25) from Mycobacterium tuberculosis H37Rv, Protein Expr.
Purif. 26 (2002) 59–64.
[36] R.G. Silva, L.P.S. Carvalho, J.S. Oliveira, C.A. Pinto, M.A. Men-
des, M.S. Palma, L.A. Basso, D.S. Santos, Cloning, overexpression
and puriWcation of functional human purine nucleoside phosphor-
ylase, Protein Expr. Purif. 27 (2003) 158–164.
[37] K.C. Kelley, K.J. Huestis, D.A. Austen, C.T. Sanderson, M.A.
Donoghue, S.K. Stikel, E.S. Kawasaki, M.S. Osburne, Regulation
of sCD4-183 gene expression from phage T7 based vectors in
Escherichia coli, Gene 156 (1995) 33–36.
[38] T.H. Grossman, E.S. Kawaski, S.R. Punreddy, M.S. Osburne,
Spontaneous cAMP-dependent derepression of gene expression in
stationary phase plays a role in recombinant expression instabil-
ity, Gene 209 (1998) 95–103.
[39] R. Bauerle, O.K. Park, Metal-catalyzed oxidation of phenylalanine-
sentitive 3-deoxy-
D-arabino-heptulosonate-7-phosphate synthase
from Escherichia coli: inactivation and destabilization by oxidation
of active-site cysteines, J. Bacteriol. 181 (1999) 1636–1642.
[40] W.T. Lowther, B.W. Matthews, Structure and function of the methio-
nine aminopeptidases, Biochim. Biophys. Acta 1477 (2000) 157–167.
[41] P.-H. Hirel, J.-M. Schmitter, P. Dessen, G. Fayat, S. Blanquet,
Extent of N-terminal methionine excision from Escherichia coli
proteins is governed by the side-chain length of the penultimate
amino acid, Proc. Natl. Acad. Sci. USA 86 (1989) 8247–8251.
[42] I.A. Shumilin, R.H. Kretsinger, R. Bauerle, Crystal structure of phen-
ylalanine-regulated 3-deoxy-
D-arabino-heptulosonate-7-phosphate
synthase from Escherichia coli, Struct. Folding Des. 7 (1999) 865–875.
[43] J.M. Ray, R. Bauerle, PuriWcation and properties of trypto-
phan-sentitive 3-deoxy-
D-arabino-heptulosonate-7-phosphate
synthase from Escherichia coli, J. Bacteriol. 173 (1991) 1894–
1901.
[44] L.R. SchoWeld, M.L. Patchett, E.J. Parker, Expression, puriWca-
tion, and characterization of 3-deoxy-
D-arabino-heptulosonate 7-
phosphate synthase from Pyrococcus furiosus, Protein Expr. Purif.
34 (2004) 17–27.
[45] K.M. Herrmann, The shikimate pathway as an entry to aromatic
secondary metabolism, Plant Physiol. 107 (1995) 7–12.
[46] C.M. Stephens, R. Bauerle, Analysis of the metal requirement of 3-
deoxy-
D-arabino-heptulosonate-7-phosphate synthase from Esch-
erichia coli, J. Biol. Chem. 266 (1991) 20810–20817.
[47] H.F. Liao, L.L. Lin, H.R. Chin, W.H. Hsu, Serine 187 is a crucial
residue for allosteric regulation of Corynebacterium glutamicum 3-
deoxy-
D-arabino-heptulosonate-7-phosphate, FEMS Microbiol.
Lett. 184 (2001) 59–64.
[48] J. Wu, D.L. Howe, R.W. Woodard, Thermotoga maritima 3-deoxy-
D-arabino-heptulosonate-7-phosphate synthase: the ancestral
eubacterial DAHP synthase?, J. Biol. Chem. 278 (2003) 27525–
27531.
[49] R.A. Jensen, E.W. Nester, Regulatory enzymes of aromatic amino
acid biosynthesis in Bacillus subtilis. I. PuriWcation and properties
of 3-deoxy-
D-arabino-heptulosonate 7-phosphate synthetase, J.
Biol. Chem. 241 (1966) 3365–3372.
[50] G. Paravicini, T. Schmidheini, G. Braus, PuriWcation and proper-
ties of 3-deoxy-
D-arabino-heptulosonate 7-phosphate synthase
(phenylalanine-inhibitable) of Saccharomyces cerevisiae, Eur. J.
Biochem. 186 (1989) 361–366.
[51] T. Wagner, I.A. Shumilin, R. Bauerle, R.H. Kretsinger, Structure
of 3-deoxy-
D-arabino-heptulosonate-7-phosphate synthase from
Escherichia coli: comparison of the Mn
2+
2-phosphoglycolate and
the Pb
2+
2-phosphoenolpyruvate complexes and implications for
catalysis, J. Mol. Biol. 301 (2000) 389–399.
[52] I.A. Shumilin, C. Zhao, R. Bauerle, R.H. Kretsinger, Allosteric
inhibition of 3-deoxy-
D-arabino-heptulosonate-7-phosphate syn-
thase alters the coordination of both substrates, J. Mol. Biol. 320
(2002) 1147–1156.
[53] I.A. Shumilin, R. Bauerle, R.H. Kretsinger, The high-resolution
structure of 3-deoxy-
D-arabino-heptulosonate-7-phosphate syn-
thase reveals a twist in plane of bound phosphoenolpyruvate, Bio-
chemistry 42 (2003) 3766–3776.
[54] R. Vicentelli, C. Bignon, A. Gruez, S. Canaan, G. Sulzenbacher, M.
Tegoni, V. Campanacci, C. Cambillau, Medium-scale structural
genomics: strategies for protein expression and crystallization,
Acc. Chem. Res. 36 (2003) 165–172.
ANEXO IV
58
ANEXO V
59
Protein Expression and PuriWcation 47 (2006) 614–620
www.elsevier.com/locate/yprep
1046-5928/$ - see front matter © 2006 Elsevier Inc. All rights reserved.
doi:10.1016/j.pep.2006.02.012
Expression, puriWcation, and circular dichroism
analysis of human CDK9
Andreia Machado Leopoldino
a
, Fernanda Canduri
b
, Hamilton Cabral
c
, Magno Junqueira
d
,
Alessandra Bernadete Trovó de Marqui
a
, Luciano H. Apponi
e
, Isabel Osório da Fonseca
f
,
Gilberto Barbosa Domont
d
, Diógenes S. Santos
f
, Sandro Valentini
e
,
Gustavo Orlando Bonilla-Rodriguez
c
, Marcelo Andrés Fossey
c
,
Walter Filgueira de Azevedo Jr.
g
, Eloiza Helena Tajara
a,¤
a
Faculdade de Medicina de São José do Rio Preto, SJRP, SP, Brazil
b
Universidade Federal do Mato Grosso do Sul, Campo Grande, MS, Brazil
c
Universidade Estadual Paulista, SJRP, SP, Brazil
d
Rede Proteômica-Rio, Universidade Federal do Rio de Janeiro, RJ, Brazil
e
Faculdade de Ciências Farmacêuticas, UNESP, Araraquara, SP, Brazil
f
Centro de Pesquisas em Biologia Molecular e Funcional/PUCRS, Porto Alegre, RS, Brazil
g
Faculdade de Biociências-PUCRS, Porto Alegre, RS, Brazil
Received 3 November 2005, and in revised form 2 February 2006
Available online 10 March 2006
Abstract
The human cyclin-dependent kinase 9 (CDK9) protein was expressed in E. coli BL21 using the pET23a vector at 30 °C. Several milli-
grams of protein were puriWed from soluble fraction using ionic exchange and ATP-aYnity chromatography. The structural quality of
recombinant CDK9 and the estimation of its secondary structure were obtained by circular dichroism. Structural models of CDK9 pre-
sented 26% of helices in agreement with the spectra by circular dichroism analysis. This is the Wrst report on human CDK9 expression in
Escherichia coli and structure analysis and provides the Wrst step for the development of CDK9 inhibitors.
© 2006 Elsevier Inc. All rights reserved.
Keywords: CDK9; Cancer; AIDS; Structure; Dichroism analysis; Molecular modeling; Expression
The cyclin-dependent kinase 9 (CDK9)
1
is a cdc2-related
kinase protein ubiquitously expressed in most cells [1] and
involved in many physiological processes, including cell
diVerentiation and apoptosis [2,3]. In contrast to other
CDKs with cell cycle regulatory functions, this serine–thre-
onine kinase exhibits protein levels unchanged in human
cells entering and progressing through the cell cycle [4].
Originally named PITALRE for its PSTAIRE-like
sequence [5,6], CDK9 is involved in transcriptional events
[7] and is regulated by the regulatory subunit, cyclin T1,
cyclin T2 or cyclin K [8,9]. Recently, a novel isoform of
CDK9 of 55 kDa [10], which originates from an alternative
upstream promoter was identiWed.
CDK9/cyclin T belongs to the multiprotein complex
P-TEFb, an elongation factor for RNA polymerase II-
directed transcription, and is responsible for the phosphor-
ylation of the C-terminal domain of the largest subunit of
RNA polymerase II [11]. It has been found that CDK9/
cyclin T1 is also required for the viral transactivator called
Tat to stimulate the processivity of RNA pol II in HIV,
*
Corresponding author. Fax: +55 17 3201 5790.
E-mail address: tajara@famerp.br (E.H. Tajara).
1
Abbreviations used: CDK9, cyclin-dependent kinase 9; LB, Luria–Ber-
tani medium; MALDI-TOF, matrix-assisted desorption ionization time-
of-Xight; CD, circular dichroism.
A.M. Leopoldino et al. / Protein Expression and PuriWcation 47 (2006) 614–620 615
suggesting a possible involvement of this kinase in AIDS
(for references, see [12]). Ammosova et al. [13] observed
dephosphorylation of CDK9 prior to its association with
HIV-1 transcription initiation complex, which might be
important for regulating HIV-1 transcription.
Recent data have indicated that Cdk9 is a critical deter-
minant of cardiac hypertrophy, in vitro and in vivo. Sano
and Schneider [14] observed that trophic signals for
increasing cardiac mass activated Cdk9, suggesting a role in
heart disease.
The CDK inhibitor Seliciclib (CYC202, R-roscovitine)
competes for the ATP binding site on the kinase. It has
greatest activity against CDK2/cyclin E, CDK7/cyclin H,
and CDK9/cyclin T. Seliciclib induces apoptosis in tumor
cell lines, reduces tumor growth in xenografts in nude mice
and is currently in phase II clinical trials for lung cancer
and B cell malignancies [15].
The functional aspects of human CDK9 have been stud-
ied by many groups but no conformational data is known.
Since the CDK9 protein is a potential therapeutic target in
AIDS, heart disease, and cancer, data on its 3D structure
becomes an essential step in the design of new and speciWc
inhibitors which could improve the treatment of patients.
In this work, we expressed and puriWed recombinant
human CDK9 in Escherichia coli (BL21) suitable for future
structural analysis. The secondary structural data obtained
using circular dichroism was in agreement with the confor-
mation obtained by homology modeling.
Materials and methods
PCR ampliWcation and cloning of human cdk9 cDNA
For RT-PCR, the RNA total was isolated from human
tissue using Trizol reagent (Invitrogen). To obtain the full-
length cdk9 cDNA, RT-PCR was carried out using the oli-
gonucleotides primers forward CgCATATggCAAAG
CAGTACGACTCGGTG and reverse gCggATCCTCAg
AAgACgCgCTCAAACTCC designed based on the cDNA
sequence from GenBank Accession No. NM_001261. The
5Ј NdeI and 3Ј BamHI restriction sites are shown in italic.
The start and stop codons are shown in bold and were
designed to be complementary to 24 and 22 bases of the 5Ј
and 3Ј ends of the CDK9 cDNA, respectively. The cdk9
cDNA (1119 bp) was ampliWed using standard PCR condi-
tions with Pfu DNA polymerase (Stratagene). To introduce
the 5Ј NdeI and 3Ј BamHI sites the ampliWed fragments
were digested with NdeI and BamHI, puriWed using the
Qiagen Gel PuriWcation, and inserted into a pET23a(+)
expression vector (Novagen), previously digested with the
same enzymes. The correct ORF was conWrmed by restric-
tion analysis and sequencing.
Expression of recombinant human CDK9
E. coli
BL21(DE3)pLys cells were transformed with the
pET23a(+)::cdk9 and recombinant colonies were selected
on LB agar plates containing carbenicillin (50 gmL
¡1
) and
chloramphenicol (50gmL
¡1
). Following transformation,
several recombinant colonies were used to inoculate 5 mL
LB medium containing carbenicillin (50 gmL
¡1
) and
chloramphenicol (50 gmL
¡1
). To determine the best con-
dition of human CDK9 protein expression, E. coli
BL21(DE3)pLys cells harboring pET23a(+)::cdk9 were
grown in the presence of IPTG (Wnal concentration of 1,
0.5, and 0.1 mM) at four diVerent temperatures (37, 30, 28,
and 20 °C) at 180 rpm after OD
600
measurements (Abs 0.4–
0.6) [16,17]. One sample was removed at each hour after
induction. The cells were harvested by centrifugation,
stored at ¡80 °C or resuspended in hypotonic lysis buVer
(10 mM Tris–HCl, pH 7.4, 25 mM NaCl, and 1 mM
EDTA), and disrupted by sonication (three short bursts,
about 10 s each, allowing the bacterial suspension to cool
on ice between each burst). The proteins and cell debris
were separated by centrifugation at 20,800g for 30 min at
4 °C. The soluble and insoluble fractions were analyzed by
SDS–PAGE.
PuriWcation of recombinant human CDK9
Preculture (20 mL) of one single colony E. coli (DE3)pLys
containing the recombinant plasmid was diluted in 500 mL
of Luria–Bertani medium (LB) supplemented with appropri-
ate antibiotics [carbenicillin (50gmL
¡1
) and chlorampheni-
col (50gmL
¡1
)]. The culture was conducted at 30 °C at
180 rpm in a shaking incubator until the cells reached mid-
log growth (OD
600
measurements of 0.4–0.6). At this point,
the expression of the target protein was induced by adding
IPTG (0.1 mM) and continued incubation at 30°C for 3 h.
The cells were harvested by centrifugation and resuspended
in 30 mL lysis buVer (10 mM Tris–HCl, pH 7.4, 25mM NaCl,
1 mM EDTA, and 1 mM PMSF). After sonication (three
short bursts, about 30 s each, allowing the bacterial suspen-
sion to cool on ice between each burst), the lysate was clari-
Wed by centrifugation for 1 h at 20,800g and 4 °C. The
supernatant was loaded over a DEAE Sepharose column
(2.5 £ 9 cm) pre-equilibrated with lysis buVer. The absorbed
proteins were eluted with a linear gradient (25–500 mM
NaCl, elution volume 60mL) by the use of a peristaltic pump
at 60 mL/h. After, the fractions were visualized in silver-
stained SDS gels, the selected CDK9 fractions were pooled
and dialyzed against a buVer containing 10 mM Hepes, pH
7.4, 25mM NaCl, and 1mM EDTA. The dialyzed CDK9
were loaded onto an ATP aYnity column (Sigma; 1 £ 2.5 cm;
2 mL) pre-equilibrated with buVer A [10mM Hepes, pH 7.4,
25 mM NaCl, 1 mM EDTA, 10% glycerol (v/v), and 0.5 mM
dithiothreitol]. After washing, bound proteins were eluted
with a 50mL linear salt gradient (25–500 mM NaCl in buVer
A). Fractions containing CDK9 were pooled and concen-
trated (up to approximately 6 mg/mL) and dialyzed using an
Amicon ultraWltration cell (MWC 10,000 Da) against 10 mM
Hepes, pH 7.4, and 1 mM EDTA. The protein content was
analyzed by SDS–PAGE and visualized using the Silver
Staining kit, Protein (GE HealthCare).
616 A.M. Leopoldino et al. / Protein Expression and PuriWcation 47 (2006) 614–620
Protein determination
The protein concentration was carried out as proposed
by Bradford [18] using bovine serum albumin as standard,
and after puriWcation by direct absorbance readings at
280 nm.
Bidimensional electrophoresis
Bidimensional electrophoresis (2DE) [19] was performed
with the IPGphor IEF system and Ruby electrophoresis
unit (GE HealthCare). After sonication, the protein sample
(100 g) was diluted in buVer containing 8 M urea, 2%
Chaps (3-[(3-cholamidopropyl)dimethylammonio]-1-pro-
panesulfonate), 0.3% DTT (dithiothreitol), 0.5% IPG buVer,
pH 3–10, bromophenol blue trace and loaded into immobi-
lized pH gradient (IPG) gel strips. Isoelectric focusing was
run at 20 °C with 8000 V for until 16,500 Vh. The IPG strips
were subsequently either stored at ¡80 °C or equilibrated
for 15 min each at room temperature in SDS equilibration
buVer (50 mM Tris–HCl, pH 8.8, 6 M urea, 2% w/v SDS,
30% glycerol, and few grains of bromophenol blue) with 1%
w/v DTT. The strips were again incubated in SDS equili-
bration buVer with 2.5% w/v iodoacetamide instead of
DTT. The strips were sealed in 12.5% SDS–PAGE and elec-
trophoresis was conducted at 30 mA/gel for 5 h. The gels
were Wxed overnight in 50% ethanol and 10% acetic acid,
washed in destain solution (50% ethanol and 5% acetic
acid) for 3 min and stained in 0.05% Coomassie blue solu-
tion in 40% methanol and 10% acetic acid for 90 min. The
gels were soaked in destain solution for 15, 45, 120, and
120 min, respectively, and preserved in 5% acetic acid. The
gels were incubated in 10% methanol for 48 h and storage
in 50% glycerol until analysis. The 2D images were scanned
by using the ImageScanner (GE HealthCare) and analyzed
by using IMAGEMASTER 2D PLATINUM software
(GE HealthCare). The intensity of each spot was Wrst pro-
cessed by background subtraction and then normalized
between gels as a proportion of the total intensity protein
from the gel. 2DE was chosen to CDK9 validation because
it separated the proteins by charge and molecular weight
(two biochemical characteristics). Therefore, it is more
eYcient for proteomics analysis and mass spectrometry
than SDS–PAGE.
MS and protein identiWcation
The matrix-assisted desorption ionization time-of-Xight
(MALDI-TOF) mass spectrometry (MS) was used to con-
Wrm the identity of the recombinant CDK9. The protein
spot of interest identiWed by 2DE absent in control, was cut
from the gel for subsequent tryptic digestion. For MALDI-
TOF MS analysis, extracted peptides were resuspended
with 0.1% triXuoroacetic acid in 50% acetonitrile, mixed
with 1 l of matrix (10 mg/ml -cyano-4-hydroxycinnamic
acid) and 1 l of this mixture was pipetted in the MALDI
plate and allowed to dry. Peptide maps were obtained using
a Voyager–DE PRO MALDI-TOF mass spectrometer
(Applied Biosystem) in positive ion reXectron mode. Pep-
tide fragment mass values were searched against the data-
base by using the MS-Fit program (http://
prospector.ucsf.edu) with the minimum parameters: Swiss-
Prot database, match with three peptides, mass tolerance of
50 ppm, and 15% coverage percentage.
Circular dichroism
Circular dichroism (CD) spectra in the UV–visible were
recorded on a Jasco J710 dichrograph operating at room
temperature, range 250–190 nm, interfaced with a PC. The
spectrum was analyzed through the self-consistent method
published by Sreerama and Wood [20], and an interactive
graphic program for calculation the secondary structures
by Deléage and Geourjon [21]. Measurements in the far UV
were carried out on 25 m cuvettes and the proteins were
dissolved in 10 mM Hepes, pH 7.4, and 1 mM EDTA.
Molecular modeling
Homology modeling is usually the method of choice
when there is a clear relationship of homology between the
sequence of a target protein and at least one known struc-
ture. Model building of the CDK9 were carried out using
the program Parmodel [22], which is a web server for auto-
mated modeling and protein structural assessment. Par-
model runs a parallelized version of MODELLER [23].
MODELLER is an implementation of an automated
approach to comparative modeling by satisfaction of spa-
tial restraints. The modeling procedure begins with align-
ment of the sequence to be modeled (target) with the
sequence of related known three-dimensional structures
(templates). This alignment is usually the input to the
program. The output is a three-dimensional model for the
target sequence containing all mainchain and sidechain
non-hydrogen atoms [23].
Two models were built, the Wrst in the absence of
ligands, and the second in the presence of ATP. The models
of CDK9 were based on the atomic coordinates of CDK2
apoenzyme (PDB code: 1HCL) [24] and CDK2:ATP (PDB
code: 1HCK) [25].
Several slightly diVerent models can be calculated by
varying the initial structure. A total of 1000 models for each
modeling were generated by the program Parmodel [22],
the Wnal models were selected based on stereochemical
quality. All optimization process was performed on a Beo-
wulf cluster with 16 nodes (BioComp, AMD Athlon XP
2100+).
Results and discussion
CDK9 is a serine–threonine kinase involved in transcrip-
tion and responsible for the activation of RNA polymerase
II. Due to its role in these processes, it is a potential thera-
peutic target in human disease. Therefore, data on its 3D
A.M. Leopoldino et al. / Protein Expression and PuriWcation 47 (2006) 614–620 617
structure are essential to the development of new and spe-
ciWc inhibitors.
Studying protein 3D structure represents a challenge in
protein biochemistry. In vitro studies such as crystallization
are reliant on the solubilization, and on appropriate quan-
tity and quality of proteins. Here, we present data that led
to the successful expression and puriWcation of human
CDK9.
Overproduction of human CDK9
CDK9 protein has been produced using pET vector
system and E. coli BL21(DE)pLys (Fig. 1, lane 4). The
expression of CDK9 at 37 °C produced insoluble pro-
teins that were resuspended in denaturing buVer con-
taining 8 M urea. However, no CD spectrum was
detected when an aliquot of dialyzed CDK9 (refolding)
was analyzed by circular dichroism at a protein concen-
tration of 13 mg/mL. Others studies [16,17] reported sim-
ilar results indicating that refolding is incomplete. The
incomplete refolding of CDK9 may be due to exposition
to urea for a long period (up to 7 days) before diluting
out the denaturant.
An optimization of the expression in diVerent tempera-
tures was used to yield soluble proteins. The best condition
was by culturing the selected colonies with the plasmid
pET23a(+)::cdk9 in LB medium at 30 °C and 0.1 mM IPTG
(Fig. 1). The expression at 37 °C resulted in high amounts of
precipitated proteins; at lower temperature (28 and 20 °C),
the expression resulted in low amounts of protein (Fig. 2).
In respect to IPTG, we found that cultures with lower
IPTG concentrations produced appropriate amounts of
soluble protein. Almost 40 mg mL
¡1
of total protein was
synthesized and the soluble proteins were recovered after
centrifugation (Table 1).
By using direct absorbance readings at 280 nm, we esti-
mated that, on average, about 10 mg of pure CDK9 was
puriWed from a 500 mL culture of transformed bacteria.
Validation and puriWcation of human CDK9
The proteins before and after induction by IPTG were
analyzed by 2DE for validation. A spot with pI and molec-
ular weight close to the theoretical values calculated for
CDK9 (pI 8.97 and 42 kDa) was present in 2DE gel after
induction but absent in control. Protein spot was excised,
in-gel digested with trypsin, and identiWed by MALDI mass
spectrometry. The processed peptide mass Wngerprint data
analyzed by the MS-Fit search identiWed the human CDK9
as the Wrst hit with coverage of 82% (data not shown).
Recombinant protein was puriWed by ionic exchange
(DEAE Sepharose) and ATP-aYnity chromatography
(Table 1; Fig. 1, lane 3). The resins and elution conditions
were chosen based on our experience acquired during
CDK2 puriWcation [27]. The CDK2 puriWcation protocol
was optimized for CDK9. The fractions of puriWed CDK9
were pooled and concentrated for future analysis and
Fig. 1. PuriWcation of the recombinant human CDK9 protein by ion
exchange and ATP-aYnity chromatography. Lane: 1, crude extract, no
CDK9 expression (control); 2, ion exchange chromatography of total sol-
uble protein; 3, ATP-aYnity chromatography of human CDK9; 4, crude
extract with recombinant human CDK9; 5, LMW protein marker (GE
HealthCare). Silver staining. The resulting CDK9 is almost 99% pure.
Fig. 2. SDS–PAGE of the crude extract containing recombinant human
CDK9. Lanes: 1–3, crude extract from expression at 20, 37, and 28 °C,
respectively. Coomassie blue staining.
Table 1
The optimal puriWcation
a
The purities were estimated from silver-stained SDS gels and the
quantiWcation was estimated using direct absorbance readings at 280 nm.
Approximate yield (%) Approximate purity
a
(%)
Cell lysate 50–60 20
DEAE column 70 50
ATP-aYnity column 90 100
618 A.M. Leopoldino et al. / Protein Expression and PuriWcation 47 (2006) 614–620
maintained at 4 °C in the presence of proteases inhibitors to
avoid degradation.
Structural analysis of CDK9
Circular dichroism is an important technique to esti-
mate secondary structure content, to validate protein
quality, and for monitoring conformational changes of
proteins due to drug binding. The CDK9 deconvoluted
CD spectrum showed that it contains 30–34% helix, 33–
36% beta, and 21–29% of coil (Fig. 3). Such CD data
permitted to obtain information about the CDK9 struc-
tures of the regions not modeled. The results suggest
that human recombinant CDK9 maintains secondary
structures and is viable to functional and structural
studies.
Fig. 4 shows the sequence alignment of CDK9 and
CDK2. The alignment presents 38% of identity, considering
the region from residues 17 to 333. The overall stereochem-
ical quality of the Wnal CDK9 apoenzyme and CDK9:ATP
models and average G-factor were calculated by the PRO-
CHECK software [26]. Analysis of the Ramachandran
plots indicates that 89.5% of the residues are in most
favored regions, and 8.4% of the residues in additional
allowed regions for the CDK9 apoenzyme whereas 87% of
the residues are in most favored regions, and 9.9% in addi-
tional allowed regions for CDK9:ATP model. The G-factor
values were 0.14 and 0.38 for CDK9 apoenzyme and
CDK9:ATP models, respectively. Ideally, scores should be
above ¡0.5.
The models for CDK9 are folded into the typical bilo-
bal structure, with the smaller N-terminal lobe consisting
predominantly of -sheet structure and the larger C-ter-
minal lobe consisting primarily of -helices. The N-termi-
nal lobe of CDK has a sheet of Wve antiparallel -strands
(1–5) and a single large helix (1). The C-terminal lobe
contains a pseudo-4-helical bundle (2, 3, 4, 6), a small -
ribbon (6–8), and two additional helices (5, 7). The
ATP molecule is found in the cleft between the two lobes.
The core (the -sheet and the helical bundle) of the CDK9
structure is very similar to that of CDK2 [25,27,28], as
shown in Figs. 5A–C.
The diVerences between the structures of CDK2 and
CDK9 are in the N-terminal (residues 1–17) and C-termi-
nal regions (residues 333–372); the CDK9 sequence is
Fig. 3. Circular dichroism spectrum of human CDK9. The protein confor-
ation consists of 30–34% helices, 33–36% beta-strands, and 21–29% coils.
Fig. 4. Sequence alignment of CDK2 with CDK9 (38% identity). The alignment was performed with the program MULTIALIGN [30].
A.M. Leopoldino et al. / Protein Expression and PuriWcation 47 (2006) 614–620 619
longer than CDK2. Therefore, the N- and C-terminal were
not modeled, due to the absence of models. These regions
can provide speciWcity to CDK9, such as the diVerences in
the loops 52–55, 88–95, 182–185, and 317-322, which can be
observed in the alignment of their primary sequences
(Fig. 4). Considering the modeled region, CDK9 shows
26% of helices. Such result is consistent with CD data of the
entire CDK9 protein.
Pinhero et al. [29] established a protocol for the expres-
sion and puriWcation of the complex His
6
-CDK9/CycT1
using recombinant baculovirus in Spodoptera frugiperda
insect cells. On an average, from 1 L culture of infected Sf9
cells, these authors could purify about 1.2 mg of pure His
6
-
CDK9/CycT1. Compared to this study, our protocol in E.
coli presents some advantages in respect to yield of puriWed
protein, fusion tags, expression system, and operating costs/
complexity.
In summary, ours results show that recombinant
human CDK9 can be cloned and expressed in E. coli and
puriWed from soluble fraction in two chromatographic
steps. Mass spectrometry identiWcation validated the
expression of CDK9. Moreover, expression at a lower
temperature using 0.1 mM IPTG resulted in a secondary
structured recombinant protein as shown by CD analysis
validating its use in future structural analysis. This is the
Wrst report on human CDK9 expression in E. coli and
structure analysis and provides the Wrst step for the devel-
opment of new inhibitors for the treatment of AIDS,
heart disease, and cancer.
Acknowledgments
This work was supported by grants from FAPESP
(Proc.01/07532-0, 02/04383-7, and 02/09388-7), FAPERJ,
CNPq, CAPES.
References
[1] X. Grana, A. De Luca, N. Sang, Y. Fu, P.P. Claudio, J. Rosenblatt,
D.O. Morgan, A. Giordano, PITALRE, a nuclear CDC2-related pro-
tein kinase that phosphorylates the retinoblastoma protein in vitro,
Proc. Natl. Acad. Sci. USA 91 (1994) 3834–3838.
[2] G. De Falco, Giordano A, CDK9: from basal transcription to cancer
and AIDS, Cancer Biol. Ther. 1 (2002) 342–347.
[3] X. Yang, M.O. Gold, D.N. Tang, D.E. Lewis, E. Aguilar-Cordova,
A.P. Rice, C.H. Herrmann, TAK, an HIV Tat-associated kinase, is a
member of the cyclin-dependent family of protein kinases and is
induced by activation of peripheral blood lymphocytes and diVerenti-
ation of promonocytic cell lines, Proc. Natl. Acad. Sci. USA 94 (1997)
12331–12336.
[4] J. Garriga, S. Bhattacharya, J. Calbo, R.M. Marshall, M. Truongcao,
D.S. Haines, X. Grana, CDK9 is constitutively expressed throughout
the cell cycle, and its steady-state expression is independent of SKP2,
Mol. Cell. Biol. 23 (2003) 5165–5173.
[5] J. Pines, The cell cycle kinases, Semin. Cancer Biol. 5 (1994) 305–313.
[6] P.D. JeVrey, A.A. Russo, K. Polyak, E. Gibbs, J. Hurwitz, J. Mass-
agué, N.P. Pavletich, Mechanism of CDK activation revealed by
the structure of a cyclinA-CDK2 complex, Nature 376 (1995)
313–320.
[7] S.H. Chao, D.H. Price, Flavopiridol inactivates P-TEFb and blocks
most RNA polymerase II transcription in vivo, J. Biol. Chem. 276
(2001) 31793–31799.
[8] J. Peng, N.F. Marshall, D.H. Price, IdentiWcation of a cyclin subunit
required for the function of Drosophila P-TEFb, J. Biol. Chem. 273
(1998) 13855–13860.
[9] T.J. Fu, J. Peng, G. Lee, D.H. Price, O. Flores, Cyclin K functions as a
CDK9 regulatory subunit and participates in RNA polymerase II
transcription, J. Biol. Chem. 274 (1999) 34527–34530.
[10] S.M. Shore, S.A. Byers, W. Maury, D.H. Price, IdentiWcation of a
novel isoform of Cdk9, Gene 307 (2003) 175–182.
[11] Y. Zhu, T. Pe’ery, J. Peng, Y. Ramanathan, N. Marshall, T. Marshall,
B. Amendt, M.B. Mathews, D.H. Price, Transcription elongation fac-
tor P-TEFb is required for HIV-1 tat transactivation in vitro, Genes
Dev. 11 (1997) 2622–2632.
[12] J. Garriga, X. Grana, Cellular control of gene expression by T-type
cyclin/CDK9 complexes, Gene 337 (2004) 15–23.
[13] T. Ammosova, K. Washington, Z. Debebe, J. Brady, S. Nekhai,
Dephosphorylation of CDK9 by protein phosphatase 2A and protein
Fig. 5. Ribbon diagram of (A) CDK2 structure; (B) CDK9 model; and (C) CDK9:ATP model. Figure was generated by MolMol [31].
620 A.M. Leopoldino et al. / Protein Expression and PuriWcation 47 (2006) 614–620
phosphatase-1 in Tat-activated HIV-1 transcription, Retrovirology 2
(2005) 47.
[14] M. Sano, M.D. Schneider, Cyclins that don’t cycle—cyclin T/cyclin-
dependent kinase-9 determines cardiac muscle cell size, Cell Cycle 2
(2003) 99–104.
[15] D.E. MacCallum, J. Melville, S. Frame, K. Watt, S. Anderson, A. Gia-
nella-Borradori, D.P. Lane, S.R. Green, Seliciclib (CYC202, R-Rosco-
vitine) induces cell death in multiple myeloma cells by inhibition of
RNA polymerase II-dependent transcription and down-regulation of
Mcl-1, Cancer Res. 65 (2005) 5399–5407.
[16] J.J. Chalmers, E. Kim, J.N. Telford, E.Y. Wong, W.C. Tacon, M.L.
Shuler, D.B. Wilson, EVects of temperature on Escherichia coli over-
producing beta-lactamase or human epidermal growth factor, Appl.
Environ. Microbiol. 56 (1990) 104–111.
[17] A. Lombardi, M. Sperandei, C. Cantale, P. Giacomini, P. GaleY,
Functional expression of a single-chain antibody speciWc for the
HER2 human oncogene in a bacterial reducing environment, Protein
Expr. Purif. 44 (2005) 10–15.
[18] M.M. Bradford, A rapid and sensitive method for the quantitation of
microgram quantities of protein utilizing the principle of protein-dye
binding, Anal. Biochem. 72 (1976) 248–254.
[19] A. Gorg, W. Postel, A. Domscheit, S. Gunther, Two-dimensional elec-
trophoresis with immobilized pH gradients of leaf proteins from bar-
ley (Hordeum vulgare): method, reproducibility and genetic aspects,
Electrophoresis 9 (1988) 681–692.
[20] N. Sreerama, R.W. Woody, Protein secondary structure from circular
dichroism spectroscopy. Combining variable selection principle and
cluster analysis with neural network, ridge regression and self-consis-
tent methods, J. Mol. Biol. 242 (1994) 497–507.
[21] G. Deleage, C. Geourjon, An interactive graphic program for
calculating the secondary structure content of proteins from cir-
cular dichroism spectrum, Comput. Appl. Biosci. 9 (1993)
197–199.
[22] H.B. Uchoa, G.E. Jorge, N.J. da Silveira, J.C. Câmera, F. Canduri,
W.F. De Azevedo, Parmodel: a web server for automated compara-
tive modeling of proteins, Biochem. Biophys. Res. Commun. 325
(2004) 1481–1486.
[23] A. Sali, T.L. Blundell, Comparative protein modelling by satisfaction
of spatial restraints, J. Mol. Biol. 234 (1993) 779–815.
[24] U. Schulze-Gahmen, J. Brandsen, H.D. Jones, D.O. Morgan, L. Mei-
jer, J. Vesely, S.-H. Kim, Multiple modes of ligand recognition: crystal
structures of cyclin-dependent protein kinase2 in complex with ATP
and two inhibitors, olomoucine and isopentenyladenine, Proteins 22
(1995) 378–391.
[25] U. Schulze-Gahmen, H.L. De Bondt, S.-H. Kim, High-resolution crys-
tal structures of human cyclin-dependent kinase 2 with and without
ATP: bound waters and natural ligand as guides for inhibitor design,
J. Med. Chem. 39 (1996) 4540–4546.
[26] R.A. Laskowski, M.W. MacArthur, D.S. Moss, J.M. Thornton, PRO-
CHECK: a program to check the stereochemical quality of protein
structures, J. Appl. Cryst. 26 (1993) 283–291.
[27] W.F.JR. De Azevedo, S. Leclerc, L. Meijer, L. Havlicek, M. Strnad, S.-
H. Kim, Inhibition of cyclin-dependent kinases by purine analogues:
crystal structure of human CDK2 complexed with roscovitine, Eur. J.
Biochem. 243 (1997) 518–526.
[28] F. Canduri, H.B. Uchoa, W.F. De Azevedo Jr., Molecular models of
cyclin-dependent kinase 1 complexed with inhibitors, Biochem. Bio-
phys. Res. Commun. 324 (2004) 661–666.
[29] R. Pinhero, P. Liaw, K. Yankulov, A uniform procedure for the puriW-
cation of CDK7/CycH/MAT1, CDK8/CycC and CDK9/CycT1, Biol.
Proced. Online 6 (2004) 163–172.
[30] F. Corpet, Multiple sequence alignment with hierarchical clustering,
Nucleic Acids Res. 16 (1988) 10881–10890.
[31] R. Koradi, M. Billeter, K. Wüthrich, MOLMOL: a program for dis-
play and analysis of macromolecular structures, J. Mol. Graph. 14
(1996) 51–55.
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