12
The Shikimate Pathway and Its Branches in Apicomplexan Parasites Craig W. Roberts, 1 Fiona Roberts, 2 Russell E. Lyons, 1 Michael J. Kirisits, 5,7 Ernest J Mui, 5,7 John Finnerty, 8 Jennifer J. Johnson, 5,7 David J. P. Ferguson, 4 John R. Coggins, 3 Tino Krell, 3 Graham H. Coombs, 3 Wilbur K. Milhous, 10 Dennis E. Kyle, 10 Saul Tzipori, 9 John Barnwell, 12 John B. Dame, 13 Jane Carlton, 11,a and Rima McLeod 5,6,7,b 1 Department of Immunology, University of Strathclyde, 2 Department of Pathology, Victoria Infirmary, and 3 Division of Biochemistry and Molecular Biology, Institute of Biomedical and Life Sciences, University of Glasgow, Glasgow, and 4 Nuffield Department of Pathology, John Radcliffe Hospital, Oxford University, Oxford, United Kingdom; 5 Departments of Ophthalmology and Visual Sciences and 6 Departments of Medicine and Pathology and Committees on Genetics and Immunology, University of Chicago, and 7 Department of Medicine, Michael Reese Hospital, Chicago, Illinois; 8 Department of Biology, Boston University, and 9 Tufts University School of Veterinary Medicine, Boston, Massachusetts; 10 Division of Experimental Therapeutics, Walter Reed Army Institute of Research and 11 National Institutes of Health, Bethesda, Maryland; 12 Department of Parasitic Diseases, Centers for Disease Control and Prevention, Chamblee, Georgia; 13 University of Florida, School of Veterinary Medicine, Gainesville The shikimate pathway is essential for production of a plethora of aromatic compounds in plants, bacteria, and fungi. Seven enzymes of the shikimate pathway catalyze sequential con- version of erythrose 4-phosphate and phosphoenol pyruvate to chorismate. Chorismate is then used as a substrate for other pathways that culminate in production of folates, ubiquinone, napthoquinones, and the aromatic amino acids tryptophan, phenylalanine, and tyrosine. The shikimate pathway is absent from animals and present in the apicomplexan parasites Toxo- plasma gondii, Plasmodium falciparum, and Cryptosporidium parvum. Inhibition of the pathway by glyphosate is effective in controlling growth of these parasites. These findings emphasize the potential benefits of developing additional effective inhibitors of the shikimate pathway. Such inhibitors may function as broad-spectrum antimicrobial agents that are effective against bac- terial and fungal pathogens and apicomplexan parasites. Overview of the Shikimate Pathway and Its Branches, Derivatives, and Organization Characterization of the shikimate pathway of apicomplexan parasites is of special interest because the pathway is absent from animals but apparently essential for survival of the para- sites [1, 2]. Thus, understanding the shikimate pathway in these organisms may assist in the development of new antiparasitic agents. The plethora of information available on the shikimate pathway of other organisms (bacteria, fungi, and plants) provides a starting point for these studies. However, existing data reveal differences in shikimate pathways of bacteria, fungi, and plants and suggest many possible modes of regulation and inhibition of the pathway. It remains to be seen whether the pathway in apicom- plexan parasites provides yet more diversity. The shikimate pathway operates in the cytosol of bacteria and fungi, but in plants it is also known to operate in plastid organelles. The pathway utilizes phosphoenol pyruvate and erythrose 4-phosphate to produce chorismate through seven catalytic steps [3, 4] (figure 1). It is a pathway with multiple branches. While branches exist from various intermediate pro- ducts of the pathway, chorismate represents the major branch point, and various branches give rise to many end products. Derivatives of chorismate include tryptophan via an anthranilate intermediate [3, 4], phenylalanine and tyrosine via prephenate or arogenate [3– 5], and vitamin K and metal chelators containing dihydroxybenzoic acid, such as enterochelin, via isochorismate [3, 4]. Chorismate also is used for production of ubiquinone and p-aminobenzoic acid (PABA), which is converted into folates. The shikimate pathway also gives rise to many secondary metabo- The Journal of Infectious Diseases 2002;185(Suppl 1):S25–36 q 2002 by the Infectious Diseases Society of America. All rights reserved. 0022-1899/2002/18504S-0004$02.00 Presented in part: Fourth International Toxoplasmosis meeting, Drymmen, Scotland, 26– 30 July 1996; Fifth International Toxoplasmosis meeting, Marshall, California, 1–6 May 1999; Keystone Symposium on Opportunistic Infections in AIDS, Keystone Colorado, 2– 8 April 1998; Woods Hole Molecular Parasitology meeting, Woods Hole, Massachusetts, September 1999; Joint Meeting of the American Society of Parasitologists and the Society of Nematologists, Monterey, California, 6– 9 July 1999. Financial support: NIH (AI-44328); Welcome Trust; Research to Prevent Blindness Foundation; Fulbright Scholar Awards; British Pathology Society Traveling Fellowship; Howard Hughes Foundation; World Health Organization; Brennan, Blackmon, Hoffman, and Koshland families; Toxo- plasmosis Research Institute. a Current affiliation: Parasite Genomics Group, Institute for Genomic Research, Rockville, Maryland. Reprints or correspondence: Dr. Rima McLeod, University of Chicago, Visual Sciences Center, AMBH S206, 5841 South Maryland, Chicago, IL 60637 ([email protected]). S25

The Shikimate Pathway and Its Branches in Apicomplexan Parasites

Embed Size (px)

Citation preview

The Shikimate Pathway and Its Branches in Apicomplexan Parasites

Craig W. Roberts,1 Fiona Roberts,2 Russell E. Lyons,1

Michael J. Kirisits,5,7 Ernest J Mui,5,7 John Finnerty,8

Jennifer J. Johnson,5,7 David J. P. Ferguson,4

John R. Coggins,3 Tino Krell,3 Graham H. Coombs,3

Wilbur K. Milhous,10 Dennis E. Kyle,10 Saul Tzipori,9

John Barnwell,12 John B. Dame,13 Jane Carlton,11,a

and Rima McLeod5,6,7,b

1Department of Immunology, University of Strathclyde, 2Department

of Pathology, Victoria Infirmary, and 3Division of Biochemistry and

Molecular Biology, Institute of Biomedical and Life Sciences,

University of Glasgow, Glasgow, and 4Nuffield Department of

Pathology, John Radcliffe Hospital, Oxford University, Oxford,

United Kingdom; 5Departments of Ophthalmology and Visual

Sciences and 6Departments of Medicine and Pathology and

Committees on Genetics and Immunology, University of Chicago, and7Department of Medicine, Michael Reese Hospital, Chicago, Illinois;

8Department of Biology, Boston University, and 9Tufts University

School of Veterinary Medicine, Boston, Massachusetts; 10Division of

Experimental Therapeutics, Walter Reed Army Institute of Research

and 11National Institutes of Health, Bethesda, Maryland;12Department of Parasitic Diseases, Centers for Disease Control

and Prevention, Chamblee, Georgia; 13University of Florida,

School of Veterinary Medicine, Gainesville

The shikimate pathway is essential for production of a plethora of aromatic compounds in

plants, bacteria, and fungi. Seven enzymes of the shikimate pathway catalyze sequential con-

version of erythrose 4-phosphate and phosphoenol pyruvate to chorismate. Chorismate is then

used as a substrate for other pathways that culminate in production of folates, ubiquinone,

napthoquinones, and the aromatic amino acids tryptophan, phenylalanine, and tyrosine. The

shikimate pathway is absent from animals and present in the apicomplexan parasites Toxo-

plasma gondii, Plasmodium falciparum, and Cryptosporidium parvum. Inhibition of the pathway

by glyphosate is effective in controlling growth of these parasites. These findings emphasize the

potential benefits of developing additional effective inhibitors of the shikimate pathway. Such

inhibitors may function as broad-spectrum antimicrobial agents that are effective against bac-

terial and fungal pathogens and apicomplexan parasites.

Overview of the Shikimate Pathway and Its Branches,

Derivatives, and Organization

Characterization of the shikimate pathway of apicomplexan

parasites is of special interest because the pathway is absent

from animals but apparently essential for survival of the para-

sites [1, 2]. Thus, understanding the shikimate pathway in these

organisms may assist in the development of new antiparasitic

agents. The plethora of information available on the shikimate

pathway of other organisms (bacteria, fungi, and plants) provides

a starting point for these studies. However, existing data reveal

differences in shikimate pathways of bacteria, fungi, and plants

and suggest many possible modes of regulation and inhibition of

the pathway. It remains to be seen whether the pathway in apicom-

plexan parasites provides yet more diversity.

The shikimate pathway operates in the cytosol of bacteria

and fungi, but in plants it is also known to operate in plastid

organelles. The pathway utilizes phosphoenol pyruvate and

erythrose 4-phosphate to produce chorismate through seven

catalytic steps [3, 4] (figure 1). It is a pathway with multiple

branches. While branches exist from various intermediate pro-

ducts of the pathway, chorismate represents the major branch

point, and various branches give rise to many end products.

Derivatives of chorismate include tryptophan via an anthranilate

intermediate [3, 4], phenylalanine and tyrosine via prephenate or

arogenate [3–5], and vitamin K and metal chelators containing

dihydroxybenzoic acid, such as enterochelin, via isochorismate

[3, 4]. Chorismate also is used for production of ubiquinone and

p-aminobenzoic acid (PABA), which is converted into folates.

The shikimate pathway also gives rise to many secondary metabo-

The Journal of Infectious Diseases 2002;185(Suppl 1):S25–36q 2002 by the Infectious Diseases Society of America. All rights reserved.0022-1899/2002/18504S-0004$02.00

Presented in part: Fourth International Toxoplasmosis meeting,Drymmen, Scotland, 26–30 July 1996; Fifth International Toxoplasmosismeeting, Marshall, California, 1–6 May 1999; Keystone Symposium onOpportunistic Infections in AIDS, Keystone Colorado, 2–8 April 1998;Woods Hole Molecular Parasitology meeting, Woods Hole, Massachusetts,September 1999; Joint Meeting of the American Society of Parasitologistsand the Society of Nematologists, Monterey, California, 6–9 July 1999.

Financial support: NIH (AI-44328); Welcome Trust; Research to PreventBlindness Foundation; Fulbright Scholar Awards; British PathologySociety Traveling Fellowship; Howard Hughes Foundation; World HealthOrganization; Brennan, Blackmon, Hoffman, and Koshland families; Toxo-plasmosis Research Institute.

a Current affiliation: Parasite Genomics Group, Institute for GenomicResearch, Rockville, Maryland.

Reprints or correspondence: Dr. Rima McLeod, University of Chicago,

Visual Sciences Center, AMBH S206, 5841 South Maryland, Chicago, IL

60637 ([email protected]).

S25

Figure 1. The shikimate pathway and major branches from chorismate. A, The shikimate pathway catalyses the conversion of phosphoenoly py-ruvate and erythrose 4-phosphate to chorismate by 7 enzymatic steps (boxed). Major branches from chorismate include ubiquinone synthesis, whichis dependent on 8 enzymes (boxed) (B); folate synthesis, which is dependent on 6 enzymes (boxed) (C); tryptophan synthesis, which is dependent on7 enzymes (boxed) (D); phenylalanine and tyrosine synthesis, which occur by separate pathways that branch from prephenate and involve the actionof a further 5 enzymes (boxed) (E). Further branches exist from chorismate in certain organisms, including isoprenoid synthesis, and from otherintermediates, such as quinate synthesis from dehydroquinate. There is currently direct evidence only for the branch leading to folates in apicomplexanparasites. * There are many variations and subtleties in tyrosine and phenylalamine synthesis in different species.

lites, including flavanones and napthoquinones [6], and there is an

alternative version of the pathway with aminated intermediates

that give rise to aminohydroxybenzoate (AHBA), the precursor of

the ansamycin antibiotics [7]. Our recent work on apicomplexans

has demonstrated a role for the shikimate pathway in the pro-

duction of PABA and folates in these organisms [1].

Twenty percent of the carbon derived from carbohydrate cat-

abolism is utilized in the shikimate pathway in certain organisms

[8]; therefore, it is a major part of their metabolism. Animals lack

the shikimate pathway enzymes and, with the exception of dihy-

drofolate reductase (DHFR), lack enzymes downstream of chor-

ismate. These observations suggested to us that this pathway

and its branches in apicomplexan parasites might provide a num-

ber of interrelated targets for antiparasitic agents [1].

The shikimate pathway is a viable herbicide target [9]. Speci-

fically, 5-enolpyruvylshikimate-3-phosphate (EPSP) synthase,

the sixth enzyme in the pathway, is the target for glyphosate, a

widely used herbicide (Roundup; Sigma) [3, 4, 9]. The com-

pound also inhibits EPSP synthase of Toxoplasma gondii and

the growth of T. gondii, Plasmodium falciparum, and Crypto-

sporidium parvum [1]. A number of other compounds that inhibit

various other enzymes in the shikimate pathway have been

described, and some inhibit the growth of bacteria [10]. It

is surprising that none of these has been developed commercially.

Two of these compounds, the fluorinated analogues of shikimate

(6R)-6-fluoroshikimate and (6S)-6-fluoroshikimate, inhibit the

growth of P. falciparum in vitro, presumably through competi-

tive inhibition of the shikimate pathway [2].

Genes encoding many shikimate pathway enzymes have been

cloned from various bacteria, plants, and fungi. In contrast, choris-

mate synthase genes of T. gondii and P. falciparum are the only

apicomplexan representatives that have been cloned and se-

quenced [1]. In addition, we have recently obtained evidence for

the presence of chorismate synthase in Eimeria bovis, Plasmo-

dium vivax, and C. parvum (unpublished data).

Our recent demonstration that the shikimate pathway is im-

portant for the survival of a number of apicomplexans has raised

a number of interesting questions that we and others are currently

addressing. First, since the shikimate pathway is generally oper-

ative in the plastid of plants [3, 4], we are interested in whether

some or all of it occurs in the plastid-like organelle of apicom-

plexans. Second, we are interested in how the apicomplexan shi-

kimate pathway is regulated, particularly as this may provide

clues regarding the best potential targets for new antimicrobial

agents. Third,what is the importance of branches of the shikimate

pathway for various stages of the life cycles of apicomplexan

parasites, including latent stages?

The Need for New Antimicrobial Agents for Treatment

of Apicomplexan Parasites

The phylum Apicomplexa consists of a large number of proto-

zoa, including species of Toxoplasma, Plasmodium, Eimeria,

Theileria, Cryptosporidium, and Babesia. These parasites are re-

sponsible for a variety of diseases in humans and livestock. For

example, T. gondii, a very common parasite, causes ocular and

neurologic lesions and can result in systemic disease and death

when acquired in utero [11]. It has been estimated to cost more

than two billion dollars per annum to care for children born with

congenital infection in the United States [12]. Moreover, toxo-

plasmosis is a devastating disease in those who are immunocom-

promised due to AIDS, chemotherapy, malignancy, transplan-

tation, or autoimmune diseases and the involved therapies [11].

The most effective chemotherapy for toxoplasmosis currently

comprises the combination of the antifolate agents pyrimeth-

amine and sulfadiazine. These are both associated with significant

adverse reactions. Pyrimethamine inhibits the action of DHFR

and thus the production of tetrahydrofolate, which is required

for a number of reactions. Since DHFR is present in humans,

pyrimethamine-induced depletion of folate cofactors can lead to

bone marrow suppression. Sulfadiazine can cause allergy and is

commonly poorly tolerated, which results in problems with com-

pliance. Furthermore, this combination therapy acts only on the

rapidly dividing tachyzoite stage associated with active disease

and does not affect the quiescent bradyzoite stage that forms

cysts throughout the body. Thus the encysted bradyzoite forms

remain a reservoir of infection even in treated individuals and

can give rise to repeated disease reactivation. This is especially

evident in congenitally infected and immunocompromised indi-

viduals. Disease reactivation is common in the eye and brain,

where the consequences can be serious. Therapy is often required

on multiple occasions and for long periods of time [11].

Malaria is caused by the apicomplexan parasites of the genus

Plasmodium. The four common species of Plasmodium (P. falci-

parum, P. malariae, P. vivax, and P. ovale) infect millions of

people and account for the death of more than 2 million children

annually. The emergence of drug resistance has limited the use-

ful life span of many antimalarial medicines. Current treatments

include inhibitors, such as pyrimethamine and sulfadoxine, of

the folate synthetic pathway [13].

C. parvum is another apicomplexan parasite that causes dis-

ease in humans and livestock. It causes debilitating and life-

threatening diarrheal illness in patients infected with human

immunodeficiency virus. C. parvum infection can also occur in

epidemics, in which it is most severe in the young and elderly.

There is currently no effective treatment for the disease caused

by this parasite [14, 15].

An advantage of developing agents that target the shikimate

pathway is their potential to inhibit the growth of bacterial

and fungal pathogens and apicomplexan parasites. There are

growing concerns associated with the emergence of multidrug-

resistant bacteria, such as Staphylococcus aureus, as well as a

number of enterococci, Streptococcus pneumoniae, and enteric

gram-negative bacteria [16]. In addition, there is now consider-

able worry about the increased occurrence of multidrug resis-

tance among Mycobacterium tuberculosis, Salmonella typhi, Shi-

Shikimate Pathway Branches in Apicomplexan ParasitesJID 2002;185 (Suppl 1) S27

gella species, and Neisseria gonorrhoeae [17]. Furthermore, in

immunocompromised patients, multiple infections with fungi

and a variety of different classes of organisms, including myco-

bacteria and apicomplexan parasites, may occur. There is thus

great need for the development of new antimicrobial agents, and

one that targeted several of these diseases would be especially

beneficial.

Evidence for the Shikimate Pathway

in Apicomplexan Parasites

Chemotherapeutic, biochemical (figures 2, 3), and genetic

(figure 4) evidence has been obtained for shikimate pathway

enzymes in apicomplexan parasites, including T. gondii, P.

falciparum, and C. parvum [1]. In addition, we have completed

sequencing the P. vivax aroC encoding chorismate synthase and

included it in the multisequence alignment in figure 4 (GenBank

accession number AF451277).

The first evidence we obtained for the presence of the shiki-

mate pathway in T. gondii was the ability of glyphosate to restrict

growth of T. gondii in vitro (figures 2A, 3). As this inhibition was

overcome by the addition of PABA to treated cultures, it is likely

that the shikimate pathway is important for supply of folate pre-

cursors in this parasite. Additional control experiments demon-

strated that the addition of PABA could reverse the inhibitory

effect of sulfadiazine. Sulfadiazine inhibits the action of dihy-

Figure 2. Functional and enzymatic evidence for the shikimate pathway in Toxoplasma gondii. A, Glyphosate inhibition of T. gondii tachyzoitegrowth in human foreskin fibroblasts (HFF) in vitro as demonstrated by [3H]uracil uptake by the parasite (3.12 mM glyphosate vs. none,P , :002). RH, T. gondii tachyzoites of the RH strain; P/S, pyrimethamine and sulfadiazine. Glyphosate toxicity for a variety of cell lines, includingHFF (dividing cells), bovine kidney cells (MDBK), and erythrocytes were assessed. The concentrations used were not toxic. B, p-aminobenzoic acid(PABA) rescue of glyphosate growth inhibition. Added PABA reversed the effects of glyphosate and sulfadiazine (S) but not pyrimethamine (P). C,Inhibition of 5-enolpyruvylshikimate-3-phosphate (EPSP) synthase enzyme activity by 1 mM glyphosate. Parasite extracts were produced at 4�C. D,In vivo activity of glyphosate and pyrimethamine against T. gondii. ND4 mice were infected intraperitoneally with T. gondii tachyzoites (RH strain)and immediately treated orally with glyphosate alone (100 mg/kg/day; no surviving mice by day 9), with pyrimethamine alone (12.5 mg/kg/day;10% surviving mice by day 9), or with these agents combined at the specified concentrations, or they were left untreated (0% survival at 9 days).Treatment of mice with either drug alone was not sufficient to prevent death (i.e., was similar to that of control mice). In contrast, 50% of mice admi-nistered glyphosate and pyrimethamine simultaneously survived infection (at 7 days after infection, P , :05). Results were similar in 2 replicateexperiments; the data shown are a composite of these experiments (10–12 mice/group). PABA and folate were administered ad libitum in the diet.Reprinted with permission from [1].

Roberts et al.S28 JID 2002;185 (Suppl 1)

dropteroate synthase and thus the incorporation of PABA into

the folate biosynthetic pathway. As expected, however, addition

of PABA to pyrimethamine-treated cultures had no antagonistic

effect on efficacy of pyrimethamine. This is because pyrimeth-

amine exerts its effect later in the pathway by inhibiting dihydro-

folate reductase (figure 2B). Confirmation of a T. gondii EPSP

synthase, which can be inhibited by glyphosate, was obtained

by measuring enzyme activity in crude T. gondii extracts

(figure 2C): 1 mM glyphosate inhibited 85% of the EPSP

synthase activity.

Ability of glyphosate to inhibit growth in vitro of asexual

erythrocytic forms of P. falciparum also was tested. At a concen-

tration of �1 mM, glyphosate completely inhibited parasite

growth, an effect that could be abrogated by treatment with

PABA or folate (figure 3A). Similarly, growth in vitro of

C. parvum was curtailed by glyphosate (figure 3B), although in

this case, addition of PABA to cultures did not abrogate inhi-

bition (data not shown). This last observation suggests that the

shikimate pathway is absolutely essential in this parasite, pre-

sumably providing precursors for biochemical pathways other

than the one that leads to folates. Alternatively, a lack of effect

from added PABA may simply reflect poor transport of the com-

pound into C. parvum. The ability of glyphosate to act in synergy

with sulfadiazine and pyrimethamine, which interfere with fo-

late synthesis by inhibiting dihydropteroate synthase and DHFR,

respectively, was tested in vitro. Glyphosate appeared to act

synergistically with both of these compounds in inhibiting in

vitro growth of T. gondii (table 1).

Glyphosate has had extensive toxicological testing and has

been defined as “practically non-toxic.” As such, it requires in

excess of 2000 mg/kg to cause acute toxicity in mammals [18].

Thus the effect of glyphosate on apicomplexan parasites was

assessed by use of a murine model of toxoplasmosis. These ex-

periments found that although glyphosate had no effect when

administered alone, it significantly reduced mortality when

given in combination with low doses of pyrimethamine (figure

2D). This demonstrates that inhibitors of EPSP synthase can act

in synergy with conventional antifolates and may be a useful

addition to the agents used against apicomplexan parasites.

This approach may have great benefit in treating drug-resistant

malaria as we also found that glyphosate is equally effective

in vitro against both pyrimethamine-sensitive and pyrimeth-

amine-resistant P. falciparum lines.

More recently, (6R) and (6S)-6-fluoroshikimate have each

been shown to inhibit the growth of P. falciparum in vitro [2].

The effect of (6R)-6-fluoroshikimate was ablated by the addition

of aromatic amino acids (phenylalanine, tyrosine, tryptophan),

PABA, and para-hydroxybenzoate (PHBA). In contrast, addi-

tion of aromatic amino acids and PHBA or of aromatic amino

acids alone was not sufficient to ablate the effects of the (6R)-6-

fluoroshikimate. The effect of PABA alone was not tested in

these experiments, making it difficult to determine if PHBA

and the aromatic amino acids are necessary to reverse the effects

of this inhibitor [2].

EPSP synthase activity is detectable in crude extracts of

T. gondii, and low levels of chorismate synthase can also be

measured in crude T. gondii extracts by use of an anaerobic

assay [1, 19]. Further studies detected 3-dehydroquinate dehy-

dratase and shikimate kinase in T. gondii lysates (3.0 and

1.6 mU/mg protein, respectively). 3-dehydroquinate dehydra-

Figure 3. Evidence for the shikimate pathway in Plasmodium falci-parum and Cryptosporidium parvum. A, Functional evidence for theshikimate pathway in P. falciparum. Glyphosate inhibition of in vitrogrowth of asexual erythrocytic forms and p-aminobenzoic acid(PABA) and folate (FA) antagonism of growth inhibition. Glyphosate(1.08 mM ) completely inhibited asexual stages of the D6 clone invitro, and both PABA and FA antagonized the inhibitory effect of gly-phosate. D6 is a clone from the African Sierra I/UNC isolate, which issensitive to chloroquine and pyrimethamine. PABA was �1000-foldmore effective than FA. B, Glyphosate inhibition of C. parvum growth.Glyphosate and paromomycin (PRM) each inhibited parasite growth(P , :001). None of the concentrations of glyphosate were toxic to themonolayer. “Media” indicates that inhibitor was not added. Reprintedwith permission from [1].

Shikimate Pathway Branches in Apicomplexan ParasitesJID 2002;185 (Suppl 1) S29

Figure 4. Comparison of deduced amino acid sequences of Toxoplasma gondii, Plasmodium falciparum, Plasmodium vivax, Synechocystis,Escherichia coli, and Solanum chorismate synthases. The function(s) of the insertions, which have no counterparts in the genes of any otherorganisms, including Synechocystis species, Solanum lycospersicum, Saccharomyces cerevisiae, Neurospora crassa, Hemophilus influenzae, orStaphylococcus aureus, remains to be determined. Pink indicates 100% identity; deep blue indicates 60% homology; light blue indicates conserva-tive substitution of residues.

tase was partially purified and shown to be heat labile, indicating

that it is similar to the type I dehydroquinate dehydratases found

in higher plants [20]. Previous studies using crude extract of

P. falciparum detected shikimate kinase (4 mU/mg protein),

3-deoxy-D-arabino-heptulosonate 7-phosphate (DAHP) synthase

(4 mU/mg protein), and shikimate dehydrogenase (90 mU/mg

protein) [21]. Thus, there is now evidence for the presence in

apicomplexan parasites of all of the shikimate pathway enzymes

except 3-dehydroquinate synthase.

Chorismate synthase genes of 2 apicomplexan parasites,

T. gondii and P. falciparum, have been identified and sequenced

(GenBank accession numbers U 93689 and AF008549, respec-

tively), and the sequence of P. vivax was recently completed

(GenBank accession number AF451277). The deduced amino

acid sequences share all the amino acids known to be highly

conserved and have considerable identity with each other and

the chorismate synthases from other species. However, both

the T. gondii and P. falciparum proteins differ from other

known chorismate synthases in possessing a number of unique

insertions (figure 4). We have also recently identified homolo-

gous genes in C. parvum and E. bovis by Southern blotting

with a probe derived from the T. gondii chorismate synthase

cDNA (Johnson J, Roberts CW, McLeod R; unpublished data).

In addition to cloning and sequencing the T. gondii and

P. falciparum chorismate synthase genes, using cDNA, we

have completed the sequencing of the T. gondii genomic aroC

(chorismate synthase) gene and have identified the P. falciparum

genomic aroC sequence on chromosome 11 in the ongoing

P. falciparum genome project (Kirisits M, Roberts CW, McLeod

R, et al.; unpublished data). The T. gondii gene has many introns,

but the P. falciparum sequence has none (Kirisits M, Roberts

CW, McLeod R, et al.; unpublished data) [22]. Of interest, the 50

non-coding region of T. gondii aroC contains a potential GCN4

binding motif (TGACTC). The possible significance of this is

discussed below in the section on Regulation of the Shikimate

Pathway. We have determined that the T. gondii chorismate

synthase gene (aroC) is single copy, which makes it amenable

to gene deletion experiments, which are in progress.

Phylogeny and Subcellular Location of the Shikimate

Pathway in Apicomplexan Parasites

Presence of an apicomplexan algal plastid-like organelle

[23–25] and nuclear-encoded enzymes translocated and func-

tioning in plant plastids (reviewed in [26]) initially provided a

conceptual basis for our work [27, 28]. Phylogenetic analysis of

the T. gondii and P. falciparum chorismate synthases, using

parsimony and distance [28] and maximum likelihood [29] anal-

yses, indicated a puzzling monophylogeny of apicomplexan and

fungal enzymes, whereas other apicomplexan nuclear genes

have been located in the clade with alveolata, which in turn is

alignedwithciliatesandplants [30],andplastidgenesandnuclear

genes targetedtotheplastidhavebeenalignedwithcyanobacteria

and algae [23, 31–33]. Thus, these phylogenetic analyses do not

provide information that is useful for determining subcellular

localization of the pathway. Fitzpatrick et al. [22] interpreted

their immunofluorescencedataasdemonstratingcytosolic locali-

zation of P. falciparum chorismate synthase. Definitive data are

needed to accurately characterize phylogeny and subcellular

localization of each of the shikimate pathway enzymes.

Potential Branches Derived from the Shikimate

Pathway in Apicomplexans

In plants, bacteria, and fungi, chorismate is used as a substrate

for the production of many end products, including ubiquinone

and the aromatic amino acids tryptophan, tyrosine, and phenyl-

alanine. The possible role of the shikimate pathway in supplying

chorismate for the synthesis of ubiquinone and aromatic amino

acids in apicomplexan parasites remains to be confirmed. It has

been demonstrated that treatment of human fibroblasts with

interferon-g induces degradation of tryptophan and thus pre-

vents the growth of intracellular T. gondii tachyzoites [34, 35].

Moreover, transfection of T. gondii tachyzoites with the

Escherichia coli trpB gene, which encodes the b subunit of

tryptophan synthase, has been shown in the presence of exoge-

nous indole to confer resistance to this effect of interferon-g

[36]. Thus, T. gondii tachyzoites would appear to be dependent

on exogenous tryptophan supplied through their host cell and

to be incapable of synthesizing this aromatic amino acid. The

Table 1. Effects of glyphosate alone and together with antimicro-bial agents that inhibit folate synthesis on replication of Toxoplasmagondii.

Experiment no.,

treatment

[3H]uracil

incorporation,

cpma

Predicted

[3H]uracil

incorporation,

cpm, for A + Bb

Ratio,

actual:predictedc

Experiment 1

Untreated 71,449 ^ 3763

Glyphosate (drug A) 28,138 ^ 2216

Sulfadiazine (drug B) 25,026 ^ 4365

Drug A + drug B 2368 ^ 418 9856 0.24

Experiment 2

Untreated 64,343 ^ 1222

Glyphosate (drug A) 25,097 ^ 1398

Pyrimethamine (drug B) 69,217 ^ 3253

Drug A + drug B 9354 ^ 2126 25,097 0.37

NOTE. Concentrations of 3.125 mM glyphosate, 6.25mg/mL21 sulfadiaz-

ine, and 0.025mg/mL21 pyrimethamine were used. These results are representa-

tive of at least 2 replicate experiments. Adapted with permission from [1].a [3H]uracil indicates incorporation of labeled tritiated uracil into T. gondii

nucleic acid; data are mean ^ SEM of a minimum of 3 replicate wells.b Predicted [3H]uracil incorporation for drug A plus drug B, if effect is only

additive not synergistic, is calculated as (cpm drug A £ cpm drug BÞ/ ðcpm of

untreated culture).c A ratio of actual:predicted of ,1 is considered synergistic; a ratio of

actual:predicted <1 is considered additive.

Shikimate Pathway Branches in Apicomplexan ParasitesJID 2002;185 (Suppl 1) S31

bradyzoites of T. gondii have not been studied in this respect, and

it is possible that they may be unable to scavenge tryptophan. In

contrast to T. gondii tachyzoites, it has been reported that trypto-

phan is not essential for the survival of P. falciparum, although

tyrosine is an essential aromatic amino acid [37]. These two

amino acids are synthesized from chorismate but by divergent

enzymatic pathways (figure 1). These observations suggest that

certain apicomplexan parasites have independently lost differ-

ent biochemical pathways from chorismate. These differences

may reflect the various evolutionary pressures operative in phy-

logenetically related parasites living in diverse environments.

Regulation of the Shikimate Pathway

There is no published information on regulation of the shiki-

mate pathway in apicomplexan parasites, but information from

the study of bacteria, fungi, and plants provides clues to how

this might occur. Understanding the regulation of the shikimate

pathway in apicomplexans may furnish valuable information

that will underpin therapeutic intervention.

Regulation of the shikimate pathway is complex and varies

greatly between organisms. The best-studied model is E. coli,

in which the shikimate pathway is regulated by controlling the

abundance and activity of DAHP synthase. Regulation occurs

at the protein level by feedback inhibition and at the transcrip-

tional level by repression [38–49]. Feedback inhibition is the

major controlling factor of these two mechanisms [38]. In Sac-

charomyces cerevisiae, aromatic amino acid biosynthesis is

under the “general control” mechanism, which enables the

organism to respond to limited amino acid supply. Thus, the

expression of genes encoding shikimate pathway enzymes aro3

and aro4, which encode DAHP synthase-Phe and -Tyr, respec-

tively, and aro1, which encodes the pentafunctional arom protein

(which catalyses steps 2–6 in the pathway), are regulated by the

control-activator protein GCN4. This protein binds to a specific

upstream promoter element (TGACTC) present in the genes it

regulates, thereby controlling their expression [50–59].

Starvation for amino acids also leads to increased synthesis of

shikimate pathway enzymes in the plant Arabidopsis thaliana

[60]. This finding shows that plants are capable of cross-pathway

metabolic regulation; however, in plants the regulation of the

pathway is complicated by the subcellular localization of the

enzymes. The bulk of the shikimate pathway enzyme activities

are located in the plastids [61], but there is also evidence for

cytosolic isoenzymes [62]. In a number of plants, such as carrot

(Dauchus carota) and A. thaliana, two isoforms of DAHP

synthase have been identified [63, 64]. For example, in tobacco

there is a plastidic Mn2+-stimulated isoform of DAHP synthase

and a Co2+-stimulated cytosolic isoform [65]. The genes encod-

ing the plastidic form of DAHP synthase in potatoes, tobacco,

tomato, and Arabidopsis species have been isolated and charac-

terized [4]. Induced expression of the plastidic form of the

enzyme has been demonstrated by wounding or pathogen attack

of potato tubers [4] and elicitor treatment or light exposure of

parsley cells [66] resulting in the accumulation of metabolites

of the shikimate pathway. Evidence suggests this induced

expression is due to increases in the levels of transcription of

the specific genes since cycloheximide and actinomycin D both

inhibit this induction [67].

Two DAHP synthase isoforms are differentially expressed in

A. thaliana, in which only one isoform (DHS1) was up-regulated

during wounding or pathogenic attack, while the other form

remained stable [64]. In tomato (Lycopersicon esculentum), two

distinct genes encoding plastidic forms of DAHP synthase have

been identified, and each has different expression levels in vari-

ous organs [68]. Other examples of tissue-specific expression

include EPSP synthase in petunia [69] and isozymes of choris-

mate synthase in tomato [70].

In the photosynthetic protozoon Euglena gracilis, two forms

of EPSP synthase have been identified. One is a monofunctional

60-kDa enzyme that appears to be localized to the chloroplast

[71], and the other is a domain of the multifunctional 165-kDa

AROM protein, which appears to be cytosolic [72]. These two

proteins are differentially expressed according to environmental

conditions, with the AROM predominant in dark-grown cells but

down-regulated in light-grown cells and with the monofunc-

tional plastid-associated form undetectable in dark-grown cells

but up-regulated by light [73].

The Shikimate Pathway and Its Branches

as Antimicrobial Agent Targets

The presence of the shikimate pathway in a large number of

pathogenic organisms, including bacteria, fungi, and apicom-

plexan parasites, and its absence from mammals make it an

attractive target for the development of antimicrobial agents.

Each enzyme in the pathway provides a potential target, and if

any of the branches (figure 1) are operative in appropriate life-

cycle stages, then this potential chemotherapeutic exploitation

is amplified even more. Subcellular organelle transit and trans-

location signals (the enzymes that cleave signal sequences) and

regulatory mechanisms for the pathway also may offer potential

for exploitation as targets of antimicrobial agents.

The shikimate pathway has already been exploited by the

commercial herbicide glyphosate, an inhibitor of EPSP syn-

thase. The presence of this pathway in a wide range of patho-

genic organisms, including bacteria, fungi, and apicomplexan

parasites, emphasizes the potential of targeting the pathway

in these organisms. Glyphosate inhibits the growth of both

gram-negative and -positive bacteria in vitro [74–77]. Certain

bacterial chorismate synthases, DHFRs, and chorismate mutases

also have been shown to be essential [1, 4].

Although relatively high concentrations of glyphosate are

required to restrict the growth of the apicomplexan parasites

Roberts et al.S32 JID 2002;185 (Suppl 1)

tested so far, the results provide impetus for the development of

new anti-shikimate pathway therapies. The relatively high con-

centration of glyphosate required raises the important question

of whether it is possible to alter the structure of glyphosate to

make better inhibitors. Extensive testing of glyphosate analogues

[78] has not provided compounds that are as effective as glypho-

sate for the inhibition of EPSP synthase. These empiric findings

ledMousdaleetal. [78] tosuggest thatglyphosatemaybeaunique

“toxophore” (i.e., that it may not be possible to improve upon the

way this small and simple molecule interacts with EPSP syn-

thase). The findings of McRobert and McConkey [79] suggest

that inhibition of chorismate synthase is lethal for P. falciparum.

The ultimate goal of inhibiting the shikimate pathway is assisted

by a number of studies that have resulted in the development of

an array of compounds, including various substrate analogues,

capable of interfering with various steps in this pathway [2].

Most of these inhibitory activities have been demonstrated by

use of assays using recombinant or purified enzymes or crude

homogenates of organisms; however, some inhibitors have been

assessed by their ability to inhibit the pathogen in vitro or in vivo.

A range of compounds inhibit EPSP synthase activity, such as

various analogues of its substrate, shikimate 3-phosphate [80–

83]. The role of these compounds as potential antimicrobials

has not been fully examined even though many of the com-

pounds were found to be active at low micromolar or even nano-

molar levels. Other enzymes in the shikimate pathway are also

amenable to inhibition. The fluorine-containing analogue of

phosphoenol pyruvate acts as a substrate for DAHP synthase,

the first enzymatic step in the shikimate pathway [84]. The maxi-

mum rate at which DAHP synthase can convert this competitive

inhibitor to the fluorine-containing analogue of DAHP is only

about 1% that for phosphoenol pyruvate. Similarly, substrate

analogues of DAHP are competitive inhibitors of dehydroqui-

nate synthase, the second enzymatic step [85, 86].

Substrate analogues of other shikimate pathway enzymes also

inhibit the shikimate pathway. Some of these analogues do not

inhibit their immediate enzyme, but are converted to substrate

analogues for subsequent enzymes in the pathway. For example,

in E. coli, (6S)-6-fluoroshikimate and (6R)-6-fluoroshikimate

are substrates for shikimate kinase and are converted to (6S)-

6-fluoro-EPSP and (6R)-6-fluoro-EPSP, respectively [10]. Where-

as (6R)-6-fluoro-EPSP is an inhibitor of chorismate synthase,

(6S)-6-fluoro-EPSP is slowly converted by chorismate syn-

thase to form 6-fluorochorismate [87, 88]. This 6-fluorochoris-

mate would appear to inhibit 4-aminobenzoic acid synthesis

and account for the ability of this compound to inhibit E. coli

in vitro [10, 89]. Of interest, 6-fluorochorismate can inhibit

chorismate synthase of Neurospora crassa [90]. The difference

observed between the E. coli and N. crassa systems may be due

to the bifunctional nature of the N. crassa enzyme and may

suggest that 6-fluoroshikimate interferes with the flavin reduc-

tase activity present on the N. crassa enzyme.

(6S)-6-fluoroshikimate protects mice from intraperitoneal

infection with E. coli, Pseudomonas aeruginosa, and S. aureus;

however, spontaneous resistance occurred at a high frequency

to this compound, preventing its development for clinical use

[10]. More recently, (6R) and (6S)-6-fluoroshikimic acid have

each been shown to inhibit the growth of P. falciparum in vitro

(IC50 of 1:5 £ 1025 and 2:7 £ 1024 M, respectively) [2]. The rela-

tively low IC50s of these compounds against P. falciparum high-

light the potential of rationally designed substrate analogues of

this pathway in the treatment of disease caused by apicomplexan

parasites.

Development of Resistance to Inhibitors

A potential concern for all new drugs is the development of

drug resistance. Although glyphosate tolerance has not been a

problem in the field, it is possible to artificially produce glypho-

sate tolerance in plants and bacteria. Glyphosate-tolerant plants

have been obtained by selecting tolerant cell lines while increas-

ing the concentration of glyphosate in growth media [91].

Glyphosate-tolerant bacterial cell lines have been produced by

mutagenesis [92]. These organisms have made it possible to

study glyphosate resistance, an important biological mechanism

with implications in medicine and agriculture.

Two different mechanisms to overcome EPSP synthase inhi-

bition by glyphosate have been identified. The first of these is

the production of glyphosate-tolerant mutant EPSP synthase. In

Salmonella typhimurium, a single amino acid substitution of

proline to serine at amino acid position 101 in EPSP synthase is

sufficient to confer resistance to glyphosate, and it is transferable

to E. coli via a cosmid [93]. In Klebsiella pneumoniae, substi-

tution of glycine-96 with alanine in EPSP synthase reduces sen-

sitivity to glyphosate [94].

The second mechanism of resistance is the production of ele-

vated levels of EPSP synthase activity. For example, in the

higher plant Corydalis sempervirens, glyphosate-resistant cells

have a 40-fold increase in EPSP synthase activity, while levels

of the other shikimate pathway enzymes remain the same as in

wild-type plants [95]. Similarly in a glyphosate-tolerant Petunia

hybridoma cell line, EPSP synthase activity is increased 15- to

20-fold [96]. This increase in activity may be the result of

increased gene expression, as in C. sempervirens [97], or by

gene amplification, as has been demonstrated in petunia [98],

carrot [99], and tobacco [100]. While the adaptive mechanism

appears to be the more common in plants and fungi, studies of

glyphosate resistance in E. gracilis have demonstrated that

either mechanism may occur [101].

Resistance to fluoroshikimate occurs spontaneously in E. coli.

This resistance correlates with a reduced ability to take up

shikimate and has been mapped at or near shiA, which is one of

at least two loci that determine shikimate uptake in E. coli

[102]. Other shikimate-like inhibitors might suffer from the

Shikimate Pathway Branches in Apicomplexan ParasitesJID 2002;185 (Suppl 1) S33

same problems, making altering modes of entry or “non-shiki-

mate”–like inhibitors more promising strategies.

Conclusions

Characterization of the shikimate pathway in apicomplexan

parasites may provide opportunities for the development of

new antimicrobial agents. The inhibitor studies that use glypho-

sate demonstrate empirically that inhibition of EPSP synthase,

and thus the shikimate pathway, is an effective means of inhibit-

ing the growth of T. gondii, P. falciparum, and C. parvum. In a

similar manner, the ability of (6R)-6-fluoroshikimate to inhibit

P. falciparum suggests that inhibition of chorismate synthesis is

another effective means for controlling the growth of at least

P. falciparum. Increasing our understanding of the shikimate

pathway in apicomplexans may not be only of biologic inter-

est—it may expedite the discovery and design of new anti-

microbial agents.

Many opportunistic pathogens, such as Pneumocystis carinii

[103], M. tuberculosis [104], C. parvum, and T. gondii simulta-

neously infect AIDS patients and rely on the shikimate pathway.

It is therefore possible that a single compound that inhibits the

shikimate pathway or combinations of synergistic compounds

that inhibit various enzyme targets in this and interrelated path-

ways [58, 105] could simultaneously eliminate multiple patho-

gens infecting a patient. Thus, they could reduce or eliminate

the problems of antimicrobial agent resistance.

A recent description of bacterial genomics and targets in the

shikimate pathway highlights and discusses the enzymes in the

shikimate and other pathways that are particularly promising tar-

gets [105]. For apicomplexans, to date, EPSP synthase, choris-

mate synthase, and DHFR are the validated targets, but no data

yet indicate a particular protein as favored, and much interesting

work remains to be accomplished.

Acknowledgments

We thank M. Gottlieb for his insights and encouragement and

E. Holfels, V. Aitchison, H. McGuinness, and K. Zawisza for their

assistance in preparing this manuscript.

References

1. Roberts F, Roberts CW, Johnson JJ, et al. Evidence for the shikimate path-

way in apicomplexan parasites. Nature 1998;393:801–5.

2. McConkey GA. Targeting the shikimate pathway in the malaria parasite

Plasmodium falciparum. Antimicrob Agents Chemother 1999;43:175–7.

3. Bentley R. The shikimate pathway—a metabolic tree with many branches.

Crit Rev Biochem Mol Biol 1990;25:307–84.

4. Herrmann K, Weaver L. The shikimate pathway. Annu Rev Plant Physiol

Plant Mol Biol 1999;50:473–503.

5. Stenmark SL, Pierson DL, Jensen RA, Glover GI. Blue-green bacteria

synthesise L-tyrosine by the pretyrosine pathway. Nature 1974;247:

290–2.

6. Dewick PM. The biosynthesis of shikimate metabolites. Nat Prod Rep

1998;15:17–58.

7. Kim C, Kirschning A, Bergon P, et al. Biosynthesis of 3-amino-5-hydroxy-

benzoic acid, the precursor of mC7N units in ansamycin antibiotics. J Am

Chem Soc 1996;118:7486–91.

8. Haslam E. Shikimic acid: metabolism and metabolites. Chichester, United

Kingdom: John Wiley & Sons, 1993.

9. Kishore GM, Shah DM. Amino acid biosynthesis inhibitors as herbicides.

Annu Rev Biochem 1988;57:627–63.

10. Davies GM, Barrett-Bee KJ, Jude DA, et al. (6S)-6-fluoroshikimic acid, an

antibacterial agent acting on the aromatic biosynthetic pathway. Anti-

microb Agents Chemother 1994;38:403–6.

11. Boyer K, McLeod R. Toxoplasmosis. In: Long S, Proeber C, Pickering L,

eds. Principles and practice of pediatric infectious diseases. New York:

Churchill Livingstone. 1998;1421–48.

12. US Department of Agriculture. Charting the costs of food safety. Food

Review. Washington, DC: USDA, May–August 1994.

13. Winstanley PA. Chemotherapy for falciparum malaria: the armoury, the

problems and the prospects. Parasitol Today 2000;16:146–53.

14. Coombs GH. Biochemical peculiarities and drug targets in Cryptospori-

dium parvum: lessons from other coccidian parasites. Parasitol Today

1999;15:333–8.

15. Tzipori S. Laboratory investigations and chemotherapy of Cryptospori-

dium parvum. In: Tzipori S, ed. Emerging energetic protozoa. New

York: Academic Press, 1999.

16. Rubinstein E. Antimicrobial resistance—pharmacological solutions. In-

fection 1999;27:S32–4.

17. Saier MHJ, Paulsen IT, Matin A. A bacterial model system for understand-

ing multi-drug resistance. Microb Drug Resist 1997;3:289–95.

18. US Department of Agriculture, Forest Service. Glyphosate pesticide fact

sheet. USDA. Available at http://infoventures.com/e-hlth/pestcide/gly-

phos.html.

19. Ramjee MK, Coggins JR, Thorneley RN. A continuous, anaerobic spectro-

photometric assay for chorismate synthase activity that utilizes photo-

reduced flavin mononucleotide. Anal Biochem 1994;220:137–41.

20. Deka RK, Anton IA, Dunbar B, Coggins JR. The characterisation of the

shikimate pathway enzyme dehydroquinase from Pisum sativum. FEBS

Lett 1994;349:397–402.

21. Dieckmann A, Jung A. Mechanisms of sulfadoxine resistance in Plasmo-

dium falciparum. Mol Biochem Parasitol 1986;19:143–7.

22. Fitzpatrick T, Ricken S, Lanzer M, Amrhein N, Macheroux P, Kappes

B. Subcellular localization and characterization of chorismate synthase

in the apicomplexan Plasmodium falciparum. Mol Microbiol 2001;

40:65–75.

23. Williamson DH, Gardner MJ, Preiser P, Moore DJ, Rangachari K, Wilson

RJ. The evolutionary origin of the 35 kb circular DNA of Plasmodium

falciparum: new evidence supports a possible rhodophyte ancestry.

Mol Gen Genet 1994;243:249–52.

24. Howe C. Plastid origin of an extrachromosomal DNA molecule from

Plasmodium. J Theor Biol 1992;158:199–205.

25. Wilson RJM, Denny P, Preiser P, et al. Complete gene map of the plastid-

like DNA of the malaria parasite Plasmodium falciparum. J Mol Biol

1996;261:155–72.

26. Haucke V, Schatz G. Import of proteins into mitochondria and chloro-

plasts. Cell Biol 1997;7:103–6.

27. McLeod R, Roberts CW, Johnson J, Roberts F, Mets L. Antimicrobial

agents, diagnostic reagents and vaccines based on unique apicomplexan

parasite metabolic pathways, enzymes and nucleic acids. Patent Coopera-

tive Treaty, 1998.

28. Roberts CW, Roberts F, Johnson JJ, et al. Plant-like metabolic pathways

provide novel antimicrobial agent targets in pathogenic, apicomplexan

parasites [abstract]. In: Keystone Symposia on Molecular and Cellular

Biology (Keystone, CO, 2–8 April 1998). 1998. Available by contacting

[email protected].

Roberts et al.S34 JID 2002;185 (Suppl 1)

29. Keeling PJ, Palmer JD, Donald RG, Roos DS, Waller RF, McFadden GI.

Shikimate pathway in apicomplexan parasites. Nature 1999;397:219–20.

30. Stokkermans TJ, Schwartzman JD, Keenan K, Morrissette NS, Tilney LG,

Roos DS. Inhibition of Toxoplasma gondii replication by dinitroaniline

herbicides. Exp Parasitol 1996;84:355–70.

31. Kohler S, Delwiche CF, Denny PW, et al. A plastid of probable green algal

origin in apicomplexan parasites. Science 1997;275:1485–9.

32. Fast NM, Kissinger JC, Roos DS, Keeling PJ. Nuclear-encoded, plastid-

targeted genes suggest a single common origin for apicomplexan and

dinoflagellate plastids. Mol Biol Evol 2001;18:418–26.

33. McFadden G, Keith M, Monholland J, Lang-Unnasch N. Plastids in human

parasites [letter]. Nature 1996;381:482.

34. Pfefferkorn ER. Interferon gamma blocks the growth of Toxoplasma gon-

dii in human fibroblasts by inducing the host cells to degrade tryptophan.

Proc Natl Acad Sci USA 1984;81:908–12.

35. Dai W, Pan H, Kwok O, Dubey JP. Human indoleamine 2,3-dioxygenase

inhibits Toxoplasma gondii growth in fibroblast cells. J Interferon Res

1994;14:313–7.

36. Sibley LD, Messina M, Niesman IR. Stable DNA transformation in

the obligate intracellular parasite Toxoplasma gondii by complemen-

tation of tryptophan auxotrophy. Proc Natl Acad Sci USA 1994;91:

5508–12.

37. Divo AA, Geary TG, Davis NL, Jensen JB. Nutritional requirements of

Plasmodium falciparum in culture. I. Exogenously supplied dialyzable

components necessary for continuous growth. J Protozool 1985;32:

59–64.

38. Ogino T, Garner C, Markley JL, Herrmann KM. Biosynthesis of aromatic

compounds: 13C NMR spectroscopy of whole Escherichia coli cells.

Proc Natl Acad Sci USA 1982;79:5828–32.

39. Doy CH, Brown KD. Control of aromatic biosynthesis: the multiplicity of

7-phospho-2-oxo-3-deoxy-D-arabino-heptonate D-erythrose-4-phosphate-

lyase (pyruvate-phosphorylating) in Escherichia coli W. Biochim Biophys

Acta 1965;104:377–89.

40. Weaver LM, Herrmann KM. Cloning of an aroF allele encoding a tyrosine-

insensitive 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase. J Bac-

teriol 1990;172:6581–4.

41. Ray JM, Yanofsky C, Bauerle R. Mutational analysis of the catalytic and

feedback sites of the tryptophan-sensitive 3-deoxy-D-arabino-heptuloso-

nate-7-phosphate synthase of Escherichia coli. J Bacteriol 1988;170:

5500–6.

42. Camakaris H, Pittard J. Regulation of tyrosine and phenylalanine

biosynthesis in Escherichia coli K-12: properties of the tyrR gene

product. J Bacteriol 1973;115:1135–44.

43. Pittard J, Wallace BJ. Distribution and function of genes concerned

with aromatic biosynthesis in Escherichia coli. J Bacteriol 1966;91:

1494–1508.

44. Garner CC, Herrmann KM. Operator mutations of the Escherichia coli

aroF gene. J Biol Chem 1985;260:3820–5.

45. Davies WD, Davidson BE. The nucleotide sequence of aroG, the gene

for 3-deoxy-D-arabinoheptulosonate-7-phosphate synthetase (phe) in

Escherichia coli K12. Nucleic Acids Res 1982;10:4045–8.

46. Zurawski G, Gunsalus RP, Brown KD, Yanofsky C. Structure and

regulation of aroH, the structural gene for the tryptophan-repressible

3-deoxy-D-arabino-heptulosonic acid-7-phosphate synthetase of Escher-

ichia coli. J Mol Biol 1981;145:47–73.

47. Lawley B, Pittard AJ. Regulation of aroL expression by TyrR protein

and Trp repressor in Escherichia coli K-12. J Bacteriol 1994;176:

6921–30.

48. DeFeyter RC, Davidson BE, Pittard J. Nucleotide sequence of the tran-

scription unit containing the aroL and aroM genes for Escherichia coli

K-12. J Bacteriol 1986;165:233–9.

49. Heatwole VM, Somerville RL. Synergism between the Trp repressor and

Tyr repressor in repression of the aroL promoter of Escherichia coli K-

12. J Bacteriol 1992;174:331–5.

50. Hinnebusch AG. Novel mechanisms of translational control in Saccharo-

myces cerevisiae. Trends Genet 1988;4:169–74.

51. Hinnebusch AG, Fink GR. Repeated DNA sequences upstream from

HIS1 also occur at several other co-regulated genes in Saccharomyces

cerevisiae and its regulation. J Biol Chem 1983;258:5238–47.

52. Duncan K, Edwards RM, Coggins JR. The Saccharomyces cerevisiae

ARO1 gene. An example of the co-ordinate regulation of five enzymes

on a single biosynthetic pathway. FEBS Lett 1988;241:83–8.

53. Arndt K, Fink GR. GCN4 protein, a positive transcription factor in yeast,

binds general control promoters at all 50 TGACTC 30 sequences. Proc

Natl Acad Sci USA 1986;83:8516–20.

54. Paravicini G, Mosch HU, Schmidheini T, Braus G. The general control

activator protein GCN4 is essential for a basal level of ARO3 gene

expression in Saccharomyces cerevisiae. Mol Cell Biol 1989;9:144–51.

55. Fukuda K, Watanabe M, Asano K, Ouchi K, Takasawa S. A mutated ARO4

gene for feedback-resistant DAHP synthase which causes both o-fluoro-

DL-phenylalanine resistance and beta-phenethyl-alcohol overproduction

in Saccharomyces cerevisiae. Curr Genet 1991;20:453–6.

56. Hinnebusch AG. Translational regulation of yeast GCN4. A window on

factors that control initiator-trna binding to the ribosome. J Biol Chem

1997;272:21661–4.

57. Struhl K. Transcriptional enhancement by acidic activators. Biochem

Biophys Acta 1996;1288:15–7.

58. Hinnebusch AG, BM Jackson, PP Mueller. Evidence for regulation of

reinitiation in translational control of GCN4 mRNA. Proc Natl Acad Sci

USA 1988;85:7279–83.

59. Hope IA, Struhl K. Functional dissection of a eukaryotic transcriptional

activator protein, GCN4 of yeast. Cell 1986;46:885–94.

60. Guyer D, Patton D, Ward E. Evidence for cross-pathway regulation of

metabolic gene expression in plants. Proc Natl Acad Sci USA 1995;

92:4997–5000.

61. Mousdale DM, Coggins JR. Subcellular localization of the common shiki-

mate-pathway enzymes in Pisum sativum. L Planta 1985;163:241–9.

62. Morris PF, Doong R, Jensen RA. Evidence from Solanum tuberosum in

support of the dual-pathway hypothesis of aromatic biosynthesis. Plant

Physiol 1989;89:10–4.

63. Suzich JA, Dean JFD, Herrmann KM. 3-deoxy-D-arabino-heptulosonate

7-phosphate synthase from carrot roots (Daucus carota) is a hysteric

enzyme. Plant Physiol 1985;75:765–70.

64. Keith B, Dong XN, Ausubel FM, Fink GR. Differential induction of

3-deoxy-D-arabino-heptulosonate 7-phosphate synthase genes in Arabi-

dopsis thaliana by wounding and pathogenic attack. Proc Natl Acad Sci

USA 1991; 88:8821–5.

65. Ganson RJ, d’Amato TA, Jensen RA. The two-isozyme system of

3-deoxy-D-arabino-heptulosonate 7-phosphate synthase in Nicotiana

silvestris and other higher plants. Plant Physiol 1986;82:203–10.

66. Henstrand JM, McCue KF, Brink K, Handa AK, Herrmann KM, Conn EE.

Light and fungal elicitor induce 3-deoxy-D-arabino-heptulosonate

7-phosphate synthase mRNA in suspension cultured cells of parsley

(Petroselinum crispum L.). Plant Physiol 1992;98:761–3.

67. McCue KF, Conn EE. Induction of 3-deoxy-D-arabino-heptulosonate-

7-phosphate synthase activity by fungal elicitor in cultures of Petro-

selinum crispum. Proc Natl Acad Sci USA 1989;86:7374–7.

68. Gorlach J, Beck A, Henstrand JM, et al. Differential expression of tomato

(Lycopersicon esculentum) genes encoding shikimate pathway isoen-

zymes. I. 3-deoxy-D-arabino-heptulosonate 7-phosphate synthase. Plant

Mol Biol 1993;23:697–706.

69. Gasser CS, Winter JA, Hironaka CM, Shah DM. Structure, expression, and

evolution of the 5-enolpyruvylshikimate-3-phosphate synthase genes of

petunia and tomato. J Biol Chem 1988;263:4280–7.

70. Gorlach J, Raesecke HR, Abel G, Wehrli R, Amrhein N, Schmid J. Organ-

Shikimate Pathway Branches in Apicomplexan ParasitesJID 2002;185 (Suppl 1) S35

specific differences in the ratio of alternatively spliced chorismate syn-

thase (LeCS2) transcripts in tomato. Plant J 1995;8:451–6.

71. Reinbothe S, Reinbothe C, Krauspe R, Parthier B. Changing gene

expression during dark-induced chloroplast dedifferentiation in Euglena

gracilis. Plant Physiol Biochem 1991;29:309–18.

72. Reinbothe S, Ortel B, Parthier B. Overproduction by gene amplification

of the multifunctional arom protein confers glyphosate tolerance to a

plastid-free mutant of Euglena gracilis. Mol Gen Genet 1993;239:

416–24.

73. Reinbothe C, Ortel B, Parthier B, Reinbothe S. Cytosolic and plastid forms

of 5-enolpyruvylshikimate-3-phosphate synthase in Euglena gracilis are

differentially expressed during light-induced chloroplast development.

Mol Gen Genet 1994;245:616–22.

74. Amrhein N, Johanning J, Schab J, Schulz A. Biochemical basis for glypho-

sate tolerance in a bacterium and plant tissue culture. FEBS Lett 1983;

157:191–6.

75. Steinrucken HC, Amrhein N. 5-enolpyruvylshikimate-3-phosphate syn-

thase of Klebsiella pneumoniae 2. Inhibition by glyphosate [N-(phospho-

nomethyl)glycine]. Eur J Biochem 1984;143:351–7.

76. Fischer RS, Berry A, Gaines CG, Jensen RA. Comparative action of

glyphosate as a trigger of energy drain in eubacteria. J Bacteriol 1986;

168:1147–54.

77. Fischer RS, Rubin JL, Gaines CG, Jensen RA. Glyphosate sensitivity of

5-enol-pyruvylshikimate-3-phosphate synthase from Bacillus subtilis

depends upon state of activation induced by monovalent cations. Arch

Biochem Biophys 1987;256:325–34.

78. Mousdale DM, Cogins JR. Amino acid synthesis. In: Kirkwood RC, ed.

Target sites for herbicide action. New York: Plenum Press, 1991:29–56.

79. McRobert L, McConkey GA. RNAi in the malaria parasite Plasmodium

falciparum [abstract 82]. In: Molecular Parasitology meeting (Woods

Hole, MA), 2000.

80. Miller MJ, Ream JE, Walker MC, Sikorski JA. Functionalized 3,5-dihy-

droxybenzoates as potent novel inhibitors of EPSP synthase. Bioorg

Med Chem 1994;2:331–8.

81. Miller MJ, Cleary DG, Ream JE, Snyder KR, Sikorski JA. New EPSP

synthase inhibitors: synthesis and evaluation of an aromatic tetrahedral

intermediate mimic containing a 3-malonate ether as a 3-phosphate surro-

gate. Bioorg Med Chem 1995;3:1685–92.

82. Marzabadi MR, Gruys KJ, Pansegrau PD, Walker MC, Yuen HK,

Sikorski JA. An EPSP synthase inhibitor joining shikimate 3-phosphate

with glyphosate: synthesis and ligand binding studies. Biochemistry

1996;35:4199–210.

83. Shah A, Font JL, Miller MJ, Ream JE, Walker MC, Sikorski JA.

New aromatic inhibitors of EPSP synthase incorporating hydroxy-

malonates as novel 3-phosphate replacements. Bioorg Med Chem 1997;

5:323–34.

84. Pilch PF, Somerville RL. Fluorine-containing analogues of intermediates

in the shikimate pathway. Biochemistry 1976;15:5315–20.

85. Bender SL, Widlanski T, Knowles JR. Dehydroquinate synthase: the use

of substrate analogues to probe the early steps of the catalyzed reaction.

Biochemistry 1989;28:7560–72.

86. Widlanski T, Bender SL, Knowles JR. Dehydroquinate synthase: the use of

substrate analogues to probe the early steps of the catalyzed reaction.

Biochemistry 1989;28:7572–82.

87. Ramjee M, Balasubramanian S, Abell C, et al. Reaction of (6R)-6-F-EPSP

with recombinant Escherichia coli chorismate synthase generates a stable

flavin mononucleotide semiquinone radical. J Am Chem Soc 1992;114:

3151–3.

88. Ramjee MN. The characterisation of and mechanistic studies on Escheri-

chia coli chorismate synthase [dissertation]. Brighton, United Kingdom:

University of Sussex, 1992.

89. Bornemann S, Ramjee MK, Balasubramanian S, et al Escherichia coli

chorismate synthase catalyzes the conversion of (6S)-6-fluoro-5-enolpy-

ruvylshikimate-3-phosphate to 6-fluorochorismate. Implications for the

enzyme mechanism and the antimicrobial action of (6S)-6-fluoroshiki-

mate. J Biol Chem 1995;270:22811–5.

90. Balasubramanian S, Davies GM, Coggins JR, Abell C. Inhibition of chor-

ismate synthase by (6R)- and (6S)-fluoro-5-enolpyruvyl shikimate

3-P04. J Am Chem Soc 1991;113:8945–6.

91. Nafziger ED, Widholm JM, Steinrucken HC, Killmer JL. Selection and

characterization of a carrot cell line tolerant to glyphosate. Plant Physiol

1984;76:571–4.

92. Comai L, Sen LC, Stalker DM. An altered aroA gene product confers

resistance to the herbicide glyphosate. Science 1983;221:370–1.

93. Stalker DM, Hiatt WR, Comai L. A single amino acid substitution in

the enzyme 5-enolpyruvylshikimate-3-phosphate synthase confers resis-

tance to the herbicide glyphosate. J Biol Chem 1985;260:4724–8.

94. Sost D, Amrhein N. Substitution of Gly-96 to Ala in the 5-enolpyruvylshi-

kimate-3-phosphate synthase of Klebsiella pneumoniae results in a

greatly reduced affinity for the herbicide glyphosate. Arch Biochem

Biophys 1990;282:433–6.

95. Smart CC, Johanning D, Muller G, Amrhein N. Selective overproduction

of 5-enol-pyruvylshikimic acid 3-phosphate synthase in a plant cell cul-

ture which tolerates high doses of the herbicide glyphosate. J Biol Chem

1985;260:16338–46.

96. Steinrucken HC, Schulz A, Amrhein N, Porter CA, Fraley RT. Over-

production of 5-enolpyruvylshikimate-3-phosphate synthase in a gly-

phosate-tolerant Petunia hybrida cell line. Arch Biochem Biophys 1986;

244:169–78.

97. Hollander-Czytko H, Johanning D, Meyer HE, Amrhein N. Molecular

basis for the overproduction of 5-enolpyruvylshikimate-3-phosphate

synthase in a glyphosate tolerant cell suspension culture of Corydalis

sempervirens. Plant Mol Biol 1988;11:215–20.

98. Shah DM, Horsch RB, Klee HJ, et al. Engineering herbicide tolerance in

transgenic plants. Science 1986;233:478–81.

99. Hauptmann RM, della Cioppa G, Smith AG, et al. Expression of gly-

phosate resistance in carrot somatic hybrid cells through the transfer of

an amplified 5-enopyruvylshikimate-3-phosphate synthase gene. Mol

Gen Genet 1988;211:357–63.

100. Goldsbrough PB, Hatch EM, Huang B, et al. Gene amplification in gly-

phosate tolerant tobacco cells. Plant Sci 1990;72:53–62.

101. Reinbothe S, Nelles A, Parthier B. N-(phosphonomethyl)glycine (glypho-

sate) tolerance in Euglena gracilis acquired by either overproduced or

resistant 5-enolpyruvylshikimate-3-phosphate synthase. Eur J Biochem

1991;198:365–73.

102. Jude DA, Ewart CDC, Thain JL, Davies GM, Nichols WW. Transport of

the antibacterial agent (6S)-6-fluoroshikimate and other shikimate ana-

logues by the shikimate transport system of Escherichia coli. Biochim

Biophys Acta 1996;1279:125–9.

103. Banerji S, Lugli EB, Miller RF, Wakefield AE. Analysis of genetic diver-

sity at the arom locus in isolates of Pneumocystis carinii. J Eukaryot

Microbiol 1995;42:675–9.

104. Garbe T, Servos S, Hawkins A, et al. The Mycobacterium tuberculosis shi-

kimate pathway genes: evolutionary relationship between biosynthetic

and catabolic 3-dehydroquinases. Mol Gen Genet 1991;228:385–92.

105. Payne DJ, Wallis NG, Gentry DR, Rosenberg M. The impact of genomics

on novel antibacterial targets. Curr Opin Drug Discov Devel 2000;

3:177–90.

Roberts et al.S36 JID 2002;185 (Suppl 1)