ArticlePDF Available

Ornithine Decarboxylase Gene Deletion Mutants of Leishmania donovani

Authors:

Abstract and Figures

A knockout strain of Leishmania donovani lacking both ornithine decarboxylase (ODC) alleles has been created by targeted gene replacement. Growth of Deltaodc cells in polyamine-deficient medium resulted in a rapid and profound depletion of cellular putrescine pools, although levels of spermidine were relatively unaffected. Concentrations of trypanothione, a spermidine conjugate, were also reduced, whereas glutathione concentrations were augmented. The Deltaodc L. donovani exhibited an auxotrophy for polyamines that could be circumvented by the addition of the naturally occurring polyamines, putrescine or spermidine, to the culture medium. Whereas putrescine supplementation restored intracellular pools of both putrescine and spermidine, exogenous spermidine was not converted back to putrescine, indicating that spermidine alone is sufficient to meet the polyamine requirement, and that L. donovani does not express the enzymatic machinery for polyamine degradation. The lack of a polyamine catabolic pathway in intact parasites was confirmed radiometrically. In addition, the Deltaodc strain could grow in medium supplemented with either 1,3-diaminopropane or 1, 5-diaminopentane (cadaverine), but polyamine auxotrophy could not be overcome by other aliphatic diamines or spermine. These data establish genetically that ODC is an essential gene in L. donovani, define the polyamine requirements of the parasite, and reveal the absence of a polyamine-degradative pathway.
Content may be subject to copyright.
Ornithine Decarboxylase Gene Deletion Mutants of
Leishmania donovani*
(Received for publication, September 25, 1998, and in revised form, November 13, 1998)
Yuqui Jiang‡§, Sigrid C. Roberts‡, Armando Jardim‡, Nicola S. Carter‡, Sarah Shih‡,
Mark Ariyanayagam
, Alan H. Fairlamb
, and Buddy Ullman‡
i
**
From the Department of Biochemistry and Molecular Biology, Oregon Health Sciences University, Portland, Oregon
97201-3098 and
Department of Biochemistry, University of Dundee, Dundee DD1 4HN, Scotland, United Kingdom
A knockout strain of Leishmania donovani lacking
both ornithine decarboxylase (ODC) alleles has been
created by targeted gene replacement. Growth of Dodc
cells in polyamine-deficient medium resulted in a rapid
and profound depletion of cellular putrescine pools, al-
though levels of spermidine were relatively unaffected.
Concentrations of trypanothione, a spermidine conju-
gate, were also reduced, whereas glutathione concentra-
tions were augmented. The Dodc L. donovani exhibited
an auxotrophy for polyamines that could be circum-
vented by the addition of the naturally occurring poly-
amines, putrescine or spermidine, to the culture me-
dium. Whereas putrescine supplementation restored
intracellular pools of both putrescine and spermidine,
exogenous spermidine was not converted back to pu-
trescine, indicating that spermidine alone is sufficient
to meet the polyamine requirement, and that L. dono-
vani does not express the enzymatic machinery for poly-
amine degradation. The lack of a polyamine catabolic
pathway in intact parasites was confirmed radiometri-
cally. In addition, the Dodc strain could grow in medium
supplemented with either 1,3-diaminopropane or 1,5-
diaminopentane (cadaverine), but polyamine auxotro-
phy could not be overcome by other aliphatic diamines
or spermine. These data establish genetically that ODC
is an essential gene in L. donovani, define the polyamine
requirements of the parasite, and reveal the absence of
a polyamine-degradative pathway.
Polyamines are cationic compounds that play essential roles
in cell proliferation, differentiation, and macromolecular syn-
thesis (1–3). Ornithine decarboxylase (ODC)
1
catalyzes the con-
version of ornithine to putrescine (1,4-diaminobutane) and is
the initial and rate-limiting enzyme in polyamine biosynthesis
in most organisms (4). The ODC enzyme of protozoan parasites
is a novel therapeutic target, because
D,L-
a
-difluoromethylor-
nithine (DFMO; eflornithine), an irreversible inhibitor of ODC
(5), exhibits notable efficacy against the central nervous system
phase of African sleeping sickness caused by Trypanosoma
brucei gambiense (3, 6). DFMO is also active against T. b.
rhodesiense and T. congolense in murine models and has
proven effective against other genera of protozoan parasites in
vivo and in vitro, including Plasmodia (7), Giardia (8), and
Leishmania (9). DFMO has been shown to induce a lethal
polyamine depletion in both T. brucei (10) and L. donovani (9),
the etiologic agent of visceral leishmaniasis, and toxicity to
both species is ameliorated by polyamine addition (3, 9).
The ability of trypanosomatids to undergo a very high fre-
quency of homologous recombination allows the disruption of
chromosomal loci with transfected drug resistance cassettes
(11, 12) and permits a direct test of gene function. This enables
the creation of conditionally lethal parasite strains whose sur-
vival and ability to propagate are dependent upon the provision
of compounds that can ameliorate the consequences of the
genetic lesion. This genetic approach is predicated on the avail-
ability of cloned genes and their flanking sequences, and mo-
lecular clones encoding ODC have been isolated and character-
ized from both T. brucei (13) and L. donovani (14). To
genetically define the role of ODC in polyamine metabolism
and to dissect the polyamine pathway in intact parasites, a null
mutant of L. donovani has been created by double-targeted
gene replacement in which both wild type ODC alleles have
been sequentially eliminated. The phenotypic dissection of the
parasites in which the ODC copy number has been genetically
altered has established the essential role of ODC in polyamine
metabolism, reveals significant discrepancies between the
polyamine pathway of the parasite and host cells, has impor-
tant implications in understanding the therapeutic relevance
of the polyamine pathway, and supports a general strategy for
the creation of attenuated strains for vaccine development in
prophylaxing leishmaniasis.
MATERIALS AND METHODS
Materials, Chemicals, and Reagents—[COOH-
14
C]Ornithine (5060
mCi/mmol) was obtained from Moravek Biochemicals (Brea, CA),
whereas [
14
C]spermidine trihydrochloride (113 mCi/mmol) and
[
14
C]spermine tetrahydrochloride (115 mCi/mmol) were procured from
Amersham Pharmacia Biotech. Diamines were purchased from Sigma.
DFMO was a gift from the Merrell Dow Research Institute (Cincinnati,
OH). The pX63-NEO and pX63-HYG plasmids used in the parasite
transfections and pSNBR used in the subcloning of a cosmid fragment
were generously provided to this laboratory by Dr. Stephen M. Beverley
(Washington University, St. Louis, MO). Purified L. donovani ODC was
furnished by Dr. Margaret A. Phillips (University of Texas Southwest-
ern Medical Center, Dallas, TX). All other materials, chemicals, and
reagents used in these experiments have been described previously
(14–16) and were of the highest quality commercially available.
Cell Culture and Preexisting Cell Lines—L. donovani promastigotes,
the extracellular insect vector form of the parasite, were grown in
DME-L, a completely defined culture medium especially designed for
the cultivation of Leishmania (17). In specified experiments, cells were
propagated in a modified DME-L medium, DME-L-CS, in which the
bovine serum albumin component of DME-L was replaced with 10%
* This work was supported in part by Grant AI 41622 from the
National Institute of Allergy and Infectious Disease and by a grant from
The Burroughs Wellcome Fund. The costs of publication of this article
were defrayed in part by the payment of page charges. This article must
therefore be hereby marked advertisement in accordance with 18
U.S.C. Section 1734 solely to indicate this fact.
§ A recipient of an N. L. Tartar Trust Fellowship from the Medical
Research Foundation of Oregon.
i
A Burroughs Wellcome Fund Scholar in Molecular Parasitology.
** To whom all correspondence should be addressed. Tel.: 503-494-
8437; Fax: 503-494-8393; E-mail: ullmanb@ohsu.edu.
1
The abbreviations used are: ODC, ornithine decarboxylase; DFMO,
a
-difluoromethylornithine; DME-L, Dulbecco’s modified Eagle-Leish-
mania medium; DME-L-CS, DME-L plus 10% chicken serum; kb, kilo-
base pairs; PCR, polymerase chain reaction.
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 274, No. 6, Issue of February 5, pp. 3781–3788, 1999
© 1999 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.
This paper is available on line at http://www.jbc.org 3781
chicken serum. The DI700 cell line is a wild type clone of the Sudanese
1S strain of L. donovani that was used for DNA isolation and library
construction and as a recipient strain in all initial transfections. For the
purposes of the genetic manipulations reported in this study, DI700 is
denoted ODC
1/1
, in which 1 refers to the wild type allele. Growth rate
determinations in the presence of polyamines and diamines were car-
ried out as described previously (9, 14).
DNA Manipulations and Library Construction—Genomic DNA was
isolated from L. donovani promastigotes by standard protocols (14, 15).
Southern blot analysis was performed as described previously (14, 15).
A cosmid library was prepared from DI700 genomic DNA that was
partially digested with Sau3A, and 3045-kb fragments were ligated
into the BamHI site of the Supercos 1 cosmid vector using protocols
described in the brochure from Stratagene (La Jolla, CA).
Isolation of a Cosmid Clone Containing ODC—A cosmid encompass-
ing ODC designated ODC-5L1 was isolated and purified using the L.
donovani ODC as a probe and stringent hybridization and wash condi-
tions as reported previously (14, 15). The L. donovani ODC was origi-
nally isolated from a bacteriophage clone as described previously by this
laboratory (14). Restriction mapping of the cosmid was performed as
described in the relevant brochure from Stratagene.
Oligonucleotide Primers—Primers used in the amplification of the 59-
and 39-flanking regions of ODC by the polymerase chain reaction (PCR)
are as follows, with their restriction enzyme sites underlined: 59F59,
59-CCCAAGCTTGTCCACGCTGCACAC-39;59F39,59-GTTGAACTGGC-
GGGCCGC-39;39F59,59-TCCCCCGGGGGATGCACGGCGACG-39; and
39F39,59-GAAGATCTGAGAGGCACTTTACTC-39.
Molecular Constructs for the Replacement of ODC Alleles—To con-
struct appropriate vectors for the replacement of each wild type ODC
allele, a 0.8-kb fragment (59F) consisting of 410 bp of 59- untranslated
DNA and 390 bp of the ODC open reading frame was ligated into the
HindIII-SalI site, and a 2.0-kb fragment of 39-untranslated DNA (39F)
was inserted into the SmaI-BglII site of both the pX63-NEO (18) and
pX63-HYG (19) vectors (Fig. 1). Both 59F and 39F were amplified by
PCR using standard reaction conditions (15 cycles; 94 °C for 60 s, 50 °C
for 30 s, 72 °C for 30 s) for the amplification of DNAs from plasmid or
cosmid DNA templates (16). 59F was generated from a 15-kb HindIII-
EcoRI fragment of ODC-5L1 encompassing ODC that had been sub-
cloned into pSNBR (20), whereas 39F was amplified from the 3.9-kb SalI
fragment described by Hanson et al. (14). Sense and antisense primers
used in the PCR were 59F59 and 59F39 for 59F and 39F59 and 39F39 for
39F, respectively. These PCR products were digested with the appropri-
ate restriction enzymes; 59F, which has an internal SalI site, was
cleaved with HindIII and SalI, whereas 39F was digested with SmaI
and BglII and ligated sequentially into pX63-NEO and pX63-HYG. The
presence of the flanking regions and their orientation in the knockout
vectors were confirmed by nucleotide sequencing (21) and restriction
mapping. The recombinant pX63-NEO and pX63-HYG vectors contain-
ing the ODC flanking regions are designated pX63-NEO-Dodc and
pX63-HYG-Dodc, respectively (Fig. 1, B and C).
Transfections—Parasites were transfected by electroporation using
conditions identical to those described previously (18). pX63-NEO-Dodc
and pX63-HYG-Dodc were linearized with HindIII and BglII and gel
purified before electroporation. In the construction of the Dodc strain,
the first wild type ODC allele was replaced with pX63-NEO-Dodc to
create the ODC/odc heterozygote (designated ODC
1/n
), whereas the
second wild type allele was deleted from the heterozygote with pX63-
HYG-Dodc to create the homozygous Dodc knockout null strain (desig-
nated ODC
n/h
). Electroporated parasites were maintained in liquid
medium for 24 h before plating on a drug-containing semisolid medium.
Drug-resistant clones transfected with pX63-NEO-Dodc were isolated
from plates containing 20
m
g/ml Geneticin (G418), whereas parasites
transfected with pX63-HYG-Dodc were selected in 20
m
g/ml G418, 50
m
g/ml hygromycin, and 100
m
M putrescine. Colonies isolated after
transfection with either pX63-NEO-Dodc or pX63-HYG-Dodc were
picked into 1.0 ml of liquid DME-L containing the relevant selective and
indispensable agents, expanded, and analyzed for the appropriate al-
lelic replacements by Southern blotting. The ODC
1/n
and ODC
n/h
trans-
formants were maintained continually in the appropriate selective me-
dia unless otherwise indicated.
ODC assay—ODC activity was measured radiometrically as de-
scribed previously (14).
Western Blotting—Polyclonal antibody to purified L. donovani ODC
was generated in rabbits by conventional methods (22). Promastigotes
were lysed by sonication, and cell supernatants were prepared by cen-
trifugation at 20,000 3 g.10
m
g of protein from each cell line were
fractionated by SDS-polyacrylamide gel electrophoresis (23), blotted
onto nitrocellulose membranes using a Semi-Dry Electrophoretic
Transfer Cell (Bio-Rad), and subjected to Western blot analysis by
standard protocols (22).
Polyamine Pool Determinations—5.0 3 10
8
parasites were harvested
and extracted for polyamine pool determinations with 10% trichloro-
acetic acid as described previously (24). An internal 1,7-diaminohep-
tane standard (40
m
M) was then added to the trichloroacetic acid super-
natants. The trichloroacetic acid was extracted with ethyl acetate as
reported previously (24), and the samples were dried on a Speed Vac
concentrator. Samples were pre-column derivatized with dansyl chlo-
ride using previously reported protocols (25), except that the sample
volumes were 100
m
l. 100
m
l of 25% proline were added to scavenge
excess dansyl chloride. The derivatized polyamines were recovered with
two ethyl acetate extractions, and the organic layers were pooled and
dried. Samples were resuspended in 200
m
l of 95% methanol/5% acetic
acid, and polyamines were analyzed by high performance liquid chro-
matography on a Beckman system equipped with a C
18
reversed phase
column (Bio-Rad) as described previously (25). Relative fluorescence
was measured on a Shimadzu RF-535 fluorescence high performance
liquid chromatography monitor at excitation and emission wavelengths
of 365 and 485 nm, respectively. Peak areas were calculated using a
Hewlett Packard HP3396 series II integrator and compared with those
of known polyamine standards.
Thiol Pool Measurements—1.0 3 10
8
cells were prepared for thiol
pool determinations and derivatized with monobromobimane as de-
scribed previously (24). Derivatized thiols were fractionated by high
performance liquid chromatography over a Vydac C
18
reversed phase
column as reported previously (26). Relative fluorescence was measured
at excitation and emission wavelengths of 395 and 480 nm, respectively.
Peak areas were calculated as described for the polyamine pool
determinations.
Radiolabeled Polyamine Incorporation Experiments—5 3 10
6
wild
type promastigotes were incubated with 2
m
Ci of [
14
C]spermidine (113
mCi/mmol) or [
14
C]spermine (115 mCi/mmol) in 5 ml of DME-L-CS
under normal growth conditions for 48 h. Polyamine pools were pro-
cessed and chromatographed as described above. 1-ml fractions were
collected, and radioactivity was quantified by liquid scintillation spec-
trometry. The positions of the radiolabeled polyamines were deter-
mined by co-injection with polyamine standards.
RESULTS
Replacement of the ODC Alleles—To disrupt the ODC locus,
a single copy gene in L. donovani (14), each allele was sequen-
tially replaced with a drug resistance cassette. The first ODC
allele was displaced with pX63-NEO-Dodc to create the ODC
1/n
heterozygote, and the second was replaced with pX63-HYG-
Dodc to create the null ODC
n/h
Dodc mutant. In each round of
transfection, 4 3 10
7
promastigotes were transfected with the
appropriate linearized DNA fragments, and ;100 drug-resis-
tant colonies were obtained from each plating. The homozy-
gotes were selected in medium supplemented with 100
m
M
putrescine, because it has been established previously that
pharmacologic simulation of a genetic deficiency of ODC in L.
donovani by the incubation of intact parasites with DFMO
could be circumvented by the addition of the diamine to the
culture medium (9). G418 was also added to the hygromycin
resistance selections to ensure that the second round of gene
targeting yielded cell lines in which pX63-HYG-Dodc had re-
placed the wild type allele of the ODC
1/n
heterozygote.
Southern blot analysis of the ODC
1/1
, ODC
1/n
, and ODC
n/h
strains revealed the new alleles that had been constituted by
homologous gene replacement events (Fig. 2). These altered
alleles could be effectively discriminated from the wild type
allele by the positions of distinct SacI sites located within the
endogenous ODC locus, pX63-NEO-Dodc, and pX63-HYG-Dodc
(Fig. 1). These allelic differences are demonstrated in Fig. 2.
Digestion of genomic DNA prepared from ODC
1/1
, ODC
1/n
,
and ODC
n/h
cells with SacI and hybridization with the 0.8-kb
59-flanking probe 59F (Fig. 1, Probe A) revealed only the ex-
pected wild type hybridization signals at 1.6 and 1.3 kb (Fig.
2A). A novel 3.2-kb band was observed in SacI-digested
genomic DNA from ODC
1/n
with a concomitant diminution of
the hybridization intensity of the 1.6-kb signal from the 59-
ODC Deficiency in L. donovani3782
flanking region of the wild type allele. A similar digestion of
ODC
n/h
DNA showed the loss of the 1.6-kb fragment from the
wild type allele and an increase in the hybridization signal at
3.2 kb. As expected from the restriction maps (Fig. 1), no
changes in the 1.3-kb signal were observed (Fig. 2A). A parallel
digestion of genomic DNA with SacI-SalI and probing with the
2.0-kb 39-flanking probe 39F (Fig. 1, Probe B) also confirmed the
presence of the new alleles in ODC
1/n
and ODC
n/h
cells and the
disappearance of the wild type counterparts. The SacI-SalI
band that hybridizes to 39F is 2.3 kb (see Fig. 1), whereas the
fragments from the alleles disrupted by pX63-NEO-Dodc and
pX63-HYG-Dodc are 2.8 kb (Fig. 2B). Finally, to establish that
the drug-resistant clones were truly deficient in ODC coding
region sequences, genomic DNA from the three genotypes was
digested with SalI and probed with a 1.3-kb BamHI-StuI frag-
ment located within the protein coding portion of ODC (Fig. 1,
Probe C). As anticipated, a 3.9-kb hybridization signal corre-
sponding to the wild type allele was observed in the ODC
1/1
line. The same fragment, exhibiting approximately half the
signal intensity of the wild type 3.9-kb band, was observed in
the ODC
1/n
heterozygote, whereas the band was absent in the
ODC
n/h
knockout. It is worth noting that after both rounds of
transfection, clones displaying genetic events other than sim-
ple gene replacements were detected by Southern blotting at a
frequency of ;20%, but these more complex genetic alterations
were not analyzed further.
ODC Expression in ODC
1/1
, ODC
1/n
, and ODC
n/h
L. dono-
vani—To evaluate the phenotypic consequences of ODC re-
placement, ODC activity and protein were measured in wild
type and genetically manipulated strains. As shown in Fig. 3,
FIG.2. Southern blot analysis of
wild type and mutant parasites. 10
m
g
of total genomic DNA from wild type
(ODC
1/1
), G418-resistant single replace-
ment (ODC
1/n
), and G418/hygromcyin-re-
sistant double replacement (ODC
n/h
) were
digested with SacI(A), SacI-SalI(B), and
SalI(C), fractionated on agarose gels, and
transferred to Nylon membranes. Nylon
membranes were then hybridized to ei-
ther 59F(A), 39F(B), or a probe to the
deleted portions of the protein coding re-
gion of ODC (C). The positions in the ODC
genomic locus to which each of these
probes correspond are indicated in Fig. 1.
The size markers in kb are indicated on
the left.
FIG.1.Restriction maps of the ODC
locus and plasmid constructs used in
targeted gene replacement protocols.
Restriction maps of the ODC locus (A) and
the pX63-NEO-Dodc (B) and pX63-HYG-
Dodc (C) knockout vectors are indicated.
The ODC and neomycin and hygromycin
phosphotransferase markers are trans-
lated from left to right and are indicated
by white unfilled rectangular boxes. The
0.8-kb 59F and 2.0-kb 39F-flanking re-
gions that were amplified by PCR and
used in Southern blotting experiments
(see Fig. 2) are indicated by the thick
black lines. Probes A, B, and C correspond
to 59F, 39F, and the ODC protein coding
region, respectively, and are indicated by
the thick gray lines. The predicted sizes of
the restriction fragments that hybridize
to ODC locus probes and all appropriate
restriction sites are shown.
ODC Deficiency in L. donovani 3783
levels of ODC activity in ODC
1/1
, ODC
1/n
, and ODC
n/h
cells
were directly proportional to the ODC copy number. ODC ac-
tivity in the heterozygous deletion mutant was ;50% that of
wild type cells, and no ODC activity could be detected in the
ODC
n/h
double knockout. Western blot analysis of ODC
1/1
,
ODC
1/n
, and ODC
n/h
cell extracts demonstrated that the ob-
served reduction in ODC activity in the single and double
knockouts was consistent with diminished cellular expression
of ODC protein (Fig. 4).
Nutritional Requirements of ODC
n/h
Parasites—The nutri-
tional requirements of the Dodc parasites were also evaluated.
As demonstrated in Fig. 5, the ODC
n/h
double knockout could
not grow in DME-L medium in the absence of putrescine,
whereas ODC
1/n
grew as proficiently as the ODC
1/1
strain
(data not shown). The addition of 200
m
M putrescine to the
culture medium averted the lethal consequences of ODC defi-
ciency, and the growth rate of ODC
n/h
cells in the presence of
putrescine was indistinguishable from that of the ODC
1/1
and
ODC
1/n
cell lines (Fig. 5). Putrescine did not affect the growth
rate of either the ODC
1/1
or ODC
1/n
cell lines. Parallel results
were obtained with all three strains cultivated in DME-L-CS
(data not shown). No morphological distinctions were noted
among the three strains by light microscopy as long as ODC
n/h
cells were maintained in putrescine-supplemented medium.
The cellular polyamine requirement in ODC
n/h
cells could
also be satisfied by the provision of 200
m
M spermidine, al-
though supplementation of the medium with equimolar sperm-
ine, a concentration that did not affect the growth of ODC
1/1
parasites, did not support the growth of the knockout (Fig. 6).
The experiments establishing whether spermidine and sperm-
ine could overcome cellular ODC deficiency were conducted in
DME-L-CS medium to avoid the polyamine toxicity attributa-
ble to the presence of minute amounts of polyamine oxidase
activity in DME-L medium. Interestingly, the polyamine re-
quirement of ODC
n/h
cells could also be fulfilled by supplement-
ing DME-L-CS with either 1,3-diaminopropane or 1,5-diamino-
pentane (cadaverine) (Fig. 6). However, the growth rate of Dodc
cells in DME-L-CS supplemented with either 1,3-diamino-
propane or cadaverine was somewhat less than that in medium
supplemented with putrescine, although the final cell densities
were similar (data not shown). The ODC
n/h
cells could be prop-
agated continuously in DME-L-CS supplemented with 200
m
M
concentrations of either putrescine, spermidine, 1,3-diamino-
propane, or cadaverine for .6 months. The Dodc parasites
could not grow in DME-L-CS supplemented with other dia-
mines, including 1,2-diaminoethane, 1,6-diaminohexane, 1,7-
diaminoheptane, 1,8-diaminooctane, 1,10-diaminodecane, and
1,12-diaminododecane, at concentrations that were nontoxic to
wild type parasites.
Polyamine and Thiol Pool Measurements—The metabolic
consequences of ODC deficiency on polyamine and thiol pools
were also evaluated in Dodc cells. Thiol pools were measured,
because Leishmania and other trypanosomatid protozoa con-
tain millimolar concentrations of the spermidine-containing
trypanothione molecule (27), a thiol that likely serves as a
general reductant in these parasites (28). As shown in Table I,
exponentially growing wild type and ODC
1/n
cells contained
commensurate concentrations of putrescine, spermidine,
trypanothione, glutathionylspermidine, and glutathione. The
FIG.3.ODC activity in wild type and mutant cells. ODC activity
was measured in extracts of wild type ODC
1/1
(), ODC
1/n
(M), and
ODC
n/h
(Œ) cells as described previously (14).
FIG.4.Western blot analysis of expressed ODC protein in wild
type and mutant parasites. Polyclonal antiserum against L. dono-
vani ODC was generated in rabbits and used to detect ODC protein in
fractionated extracts obtained from exponentially growing ODC
1/1
,
ODC
1/n
, and ODC
n/h
cells. Molecular mass markers are indicated on the
right.
FIG.5.Growth of wild type and mutant parasites in DME-L-CS
medium in the absence or presence of putrescine. The ability of
ODC
1/1
() and ODC
n/h
(f, M) cells to grow in DME-L-CS in the
absence (empty symbols) or presence (filled symbols) of 200
m
M putres-
cine is indicated. This experiment is representative of approximately
four others, each of which yielded similar results.
ODC Deficiency in L. donovani3784
intracellular putrescine and spermidine pools of the Dodc cells
expanded in putrescine-supplemented medium were also com-
parable to those of wild type and ODC
1/n
cells, although the
thiol concentrations were significantly higher. No spermine
was detected in any of these cell lines, which is consistent with
previous observations that L. donovani lack spermine (9).
The removal of Dodc cells from putrescine-supplemented
DME-L-CS precipitated a rapid depletion of cellular putrescine
pools, although the levels of spermidine remained relatively
constant after 12 days of maintenance in medium lacking pu-
trescine (Fig. 7A). Reduction in ODC
n/h
cell numbers was not
observed until day 7 after removal of the exogenous putrescine,
and some augmentation in the cell density was observed
through the first 4 days of incubation (Fig. 7). This marginal
cell proliferation could be attributed to the fact that Leishma-
nia accommodate sufficient polyamine pools to maintain via-
bility and enable minimal growth in the absence of both poly-
amine biosynthesis and polyamine salvage. However,
throughout the incubation period in unsupplemented DME-L-
CS, levels of trypanothione in Dodc cells were much lower than
those observed in ODC
1/1
, ODC
1/n
, or ODC
n/h
parasites grown
in putrescine-containing medium, although trypanothione lev-
els remained relatively constant, albeit low, after 24 h in the
absence of putrescine (Fig. 7B). Glutathionylspermidine levels
fluctuated slightly in the Dodc cell line maintained in medium
lacking polyamine, whereas cellular glutathione concentra-
tions increased fairly substantially, i.e. ;2-fold.
Polyamines and their conjugates were also measured in
ODC
n/h
cells that had been propagated for .3 weeks in pu-
trescine-deficient DME-L-CS supplemented with spermidine.
Under these conditions, putrescine concentrations in the Dodc
strain were 4% of those of wild type parasites, whereas sper-
midine and trypanothione pools were comparable. Polyamine
and thiol pools were also measured in Dodc parasites grown in
either 1,3-diaminopropane or cadaverine (Table I). Both dia-
mines were accumulated by the ODC
n/h
cells, and a concomi-
tant depletion of naturally occurring polyamines was observed.
Cellular putrescine pools were negligible, and spermidine pools
were markedly depleted in Dodc cells maintained in either
1,3-diaminopropane or cadaverine compared with the knockout
parasites grown in putrescine or wild type parasites. The null
mutant grown in either spermidine, spermine, or cadaverine
also appeared to accumulate small amounts of 1,3-diaminopro-
pane. The origin of this anomaly could be imputed to contam-
inants in the spermidine, spermine, and cadaverine additives.
Trypanothione pools in ODC
n/h
cells grown in DME-L-CS sup-
plemented with either 1,3-diaminopropane or cadaverine were
also very low (2.0 and 10.5% of the levels found in putrescine-
supplemented Dodc parasites; Table I).
Polyamine Catabolism in L. donovani—The failure of sper-
midine supplementation to augment cellular putrescine pools
of Dodc parasites suggested that L. donovani lack a polyamine-
degradative pathway. To verify this conjecture, wild type par-
asites were incubated with either [
14
C]spermidine or [
14
C]sper-
mine to determine whether the polyamines could be
catabolized. As demonstrated in Fig. 8, L. donovani do not
convert extracellular spermidine or spermine to putrescine,
although each radiolabel is accumulated by the parasites in
unaltered form. Intracellular [
14
C]spermidine is observed in
promastigotes that had been incubated with [
14
C]spermine, but
this could be ascribed to a trace [
14
C]spermidine contaminant
present in the radiolabeled spermine (data not shown).
DISCUSSION
The creation of a Dodc strain of Leishmania by double-tar-
geted gene replacement establishes that ODC plays an essen-
tial housekeeping function in this genus of protozoan parasite.
Using independent drug resistance cassettes, ODC coding se-
quences were expunged from wild type L. donovani in two
discrete steps, and the presumptive heterozygote generated
after the first round of transfection and selection appeared to
contain only one intact ODC allele as measured by loss of the
hybridization signal and a ;50% reduction of ODC activity and
protein as compared with the ODC
1/1
parent. These data con-
firmed previous results obtained by Southern blotting that L.
donovani is diploid at the ODC locus, and that ODC is a single
copy gene (14). The observation that Dodc cells cannot survive
without putrescine supplementation of the culture medium
demonstrates genetically that ODC and polyamines are indis-
pensable to L. donovani and underscores the potential of the
FIG.6.Growth of Dodc mutants in DME-L-CS medium supple-
mented with various polyamines and diamines. The ability of
ODC
n/h
cells to grow in DME-L-CS in the presence of 200
m
M concen-
trations of putrescine (), spermidine (f), spermine (), 1,3-diamino-
propane (E), or cadaverine (M) is compared.
T
ABLE I
Polyamine and thiol levels in wild type and mutant L. donovani
Polyamine and thiol levels (nmol/10
8
cells) were determined as described under “Materials and Methods” for wild type (ODC
1/1
), heterozygous
(ODC
1/h
), and homozygous (ODC
n/h
). These pools were measured in exponentially growing cells in DME-L-CS in triplicate with the indicated
additions. GSH, glutathione; GS-SPD, glutathionylspermidine; T(SH)2, trypanothione; PUT, putrescine; SPD, spermidine; CAD, cadaverine; 1,3
DP (1,3-diaminopropane); SPM, spermine; ND, not detected.
Cell line Additions GSH GS-SPD T(SH)2 PUT SPD CAD 1,3 DP SPM
ODC
1/1
None 2.15 6 0.37 0.33 6 0.07 8.23 6 1.08 5.00 6 0.06 2.60 6 0.06 ND ND ND
ODC
1/h
None 3.31 6 0.92 0.17 6 0.08 5.31 6 0.43 4.92 6 0.44 2.80 6 0.20 ND ND ND
ODC
n/h
1,3-Diaminopropane 4.41 6 1.00 0.01 6 0.01 0.26 6 0.08 0.06 6 0.04 0.30 6 0.04 ND 19.80 6 0.24 ND
ODC
n/h
Putrescine 9.81 6 1.80 1.23 6 0.41 12.86 6 2.75 3.07 6 0.74 3.14 6 0.34 ND ND ND
ODC
n/h
Cadaverine 2.87 6 0.22 0.05 6 0.03 1.36 6 0.23 0.04 6 0.02 0.40 6 0.06 4.66 6 0.48 0.11 6 0.06 ND
ODC
n/h
Spermidine 6.91 6 1.23 0.54 6 0.10 7.04 6 0.79 0.19 6 0.08 3.02 6 0.32 ND 0.46 6 0.03 ND
ODC
n/h
Spermine 7.50 6 3.10 0.33 6 0.21 3.09 6 0.96 0.76 6 0.02 2.42 6 0.16 ND 1.60 6 0.10 0.36 6 0.02
ODC Deficiency in L. donovani 3785
polyamine pathway as a therapeutic target. The requirement
for ODC is consistent with previous observations that DFMO
toxicity in L. donovani can be bypassed by the addition of
putrescine to the culture medium (9). Furthermore, the growth
phenotype of the Dodc strain authenticates that ODC is the sole
enzyme that initiates polyamine biosynthesis in L. donovani.
This conclusion is an important distinction, because Esche-
richia coli and plants can synthesize putrescine from arginine
via an arginine decarboxylase-agmatine ureohydrolase path-
way (2). Moreover, there have been reports (29, 30), although
unconfirmed (31, 32), that T. cruzi, a protozoan parasite related
to Leishmania, expresses an arginine decarboxylase activity
that can be targeted by specific inhibitors.
Genetic studies have also demonstrated that ODC is indis-
pensable for long-term proliferation and viability of T. brucei
(33). After replacement of both ODC alleles with drug resist-
ance cassettes, ODC-deficient T. brucei incubated without poly-
amines growth arrested at the G
1
-S-phase transition of the cell
cycle. Interestingly, a small percentage of the arrested Dodc
population remained viable 7–8 weeks after putrescine with-
drawal (33). ODC-deficient mutants of mammalian cells (34)
and Neurospora crassa (35) also could not thrive in the absence
of polyamine supplement, confirming that ODC activity is
mandatory for the growth of these cells. However, ODC is not
essential for E. coli (36), which has an alternative pathway for
polyamine synthesis, or for some mutant strains of Saccharo-
myces cerevisiae that are both ODC and polyamine deficient
(37). Moreover, recent studies have strongly implied that T.
cruzi lack a polyamine biosynthetic pathway altogether and are
therefore obligate scavengers of polyamines (31, 32).
Although the present results establish a strict requirement
for polyamines in L. donovani, the precise mechanism by which
ODC deficiency triggers lethality has not been definitively es-
tablished. Removal of Dodc cells from putrescine-supplemented
medium triggered a rapid and virtually complete depletion of
cellular putrescine pools without a concomitant diminution in
spermidine levels. A similar obliteration of putrescine pools is
also observed when wild type L. donovani are treated with
DFMO (9). Incubation of ODC
n/h
parasites in the absence of
exogenous polyamines also prompted a rapid decrease in the
FIG.7.Polyamine and thiol pools in ODC
n/h
cells incubated in
medium lacking polyamine. ODC
n/h
cells were removed from DME-L
supplemented with 200
m
M putrescine and resuspended in fresh DME-L
lacking putrescine as described under “Materials and Methods.” At the
time of resuspension and at 24-h intervals thereafter, 5 3 10
8
cells were
removed for putrescine () and spermidine (E) measurements (A), and
10
8
cells were removed for the determination of trypanothione (f),
glutathionylspermidine (M), and glutathione (Œ) pools (B). Cell density
was monitored by Coulter counting ().
FIG.8.Polyamine catabolism in L. donovani. ODC
1/1
L. dono-
vani were incubated with [
14
C]spermidine (A)or[
14
C]spermine (B)as
reported under “Materials and Methods.” Cells were harvested and
washed, and polyamines were fractionated as described. 1.0-ml frac-
tions were then counted by liquid scintillation. The migration of known
standards of putrescine (PUT), spermidine (SPD), and spermine (SPN)
is indicated.
ODC Deficiency in L. donovani3786
cellular levels of trypanothione, which were subsequently
maintained at a low level, whereas concentrations of glutathi-
one increased steadily throughout the prolonged incubation.
The depletion of trypanothione could be ascribed either to the
reduced flux through the polyamine biosynthetic pathway or to
trypanothione turnover to replenish spermidine pools. It is
possible that this augmentation of glutathione is a cellular
compensation mechanism to preserve the reductive potential of
the intracellular environment when trypanothione pools are
reduced. The trypanothione depletion and glutathione eleva-
tion in putrescine-depleted ODC
n/h
L. donovani parallels re-
sults obtained with T. brucei in which glutathionylspermidine
and trypanothione levels were both substantially depleted and
glutathione levels were augmented after DFMO treatment
(26). Thus, cessation of parasite growth in polyamine-deficient
medium correlates with depletion of both the putrescine and
trypanothione pools.
Circumvention of the conditionally lethal Dodc mutation can
be achieved by the addition of either putrescine or spermidine
to the culture medium. Given that the null mutant propagated
in spermidine-supplemented medium exhibited a profound re-
duction of the putrescine pool, the most obvious inference is
that L. donovani does not require intracellular putrescine to
survive, and that spermidine is both sufficient and necessary to
satisfy the polyamine requirement of the parasite. Spermine, a
major polyamine of mammalian cells, is not detected in L.
donovani (9), supporting the lack of a spermine synthase ac-
tivity. Moreover, exogenous spermine does not satisfy the poly-
amine requirement of the Dodc promastigotes, a distinction
from what is observed in mammalian cells in which exogenous
spermine can circumvent a genetic deficiency in ODC activity
(34). In mammalian cells, spermine is converted to spermidine
and spermidine is converted to putrescine by an identical two-
enzyme pathway (38) consisting of spermidine/spermine N
1
-
acetyltransferase and polyamine oxidase. The inability of Dodc
cells to use spermine as a polyamine source or to replenish
their putrescine pools from extracellular spermidine implied
that L. donovani lacks the spermidine/spermine N
1
-acetyl-
transferase/polyamine oxidase pathway altogether. This was
then confirmed radiometrically using both [
14
C]spermidine and
[
14
C]spermine. Considering that parasitic nematodes (39) and
E. coli (40) can acetylate naturally occurring polyamines, the
lack of a counterpart pathway in L. donovani is unusual. Thus,
it seems that the equilibrium between the putrescine and sper-
midine pools in L. donovani is governed exclusively by spermi-
dine synthase activity, unlike mammalian cells, in which in-
tracellular polyamines are regulated by a delicate balance
among the anabolic enzymes, spermidine and spermine syn-
thase, and the catabolic enzymes, spermidine/spermine N
1
-
acetyltransferase and polyamine oxidase. Although the poly-
amine aminopropyltransferases have not been generally
embraced as therapeutic targets, the limited complement of
polyamine enzymes in L. donovani provides a rational basis for
the utilization of spermidine synthase inhibitors (41) in con-
junction with exogenous spermidine/spermine in therapies.
The polyamine requirement of Dodc cells can also be fulfilled
by either 1,3-diaminopropane or cadaverine, because the poly-
amine auxotrophs can grow continuously in DME-L-CS supple-
mented with either diamine, and each is taken up and accu-
mulated by the parasites. Essentially no putrescine was
detected in the knockout cells incubated with either diamine,
which supports the contention that putrescine is not an essen-
tial metabolite for L. donovani. Small quantities of spermidine,
however, were observed in the 1,3-diaminopropane- and cadav-
erine-propagated cultures. The source of this spermidine is
unclear, although it may be present in the supplemented cul-
ture medium in insufficient amounts to support the growth of
the Dodc cells. Glutathionylspermidine and trypanothione lev-
els were also markedly reduced compared with either wild type
parasites or Dodc cells propagated in putrescine-supplemented
medium. Although previous investigators detected significant
amounts of homotrypanothione, the cadaverine-containing an-
alog of trypanothione, in T. cruzi (31), we were unable to
distinguish homotrypanothione and trypanothione in our high
performance liquid chromatography system. These data imply
that these thiols are not essential for the continual propagation
of L. donovani promastigotes, at least in the absence of envi-
ronmental insult.
The ability to create polyamine auxotrophs of Leishmania by
deletion of the ODC locus suggests that ODC could be targeted
by live parasite vaccination strategies against leishmaniasis. A
similar vaccine-based approach has been inaugurated using
thymidine auxotrophs of L. major in which the dihydrofolate
reductase-thymidylate synthase locus has been replaced. These
dihydrofolate reductase-thymidylate synthase null mutants in-
duce protective immunity in mice and do not precipitate dis-
ease (42). We are currently evaluating the ability of our Dodc to
infect and proliferate within human macrophages, the cell type
in which the human stage of the parasite resides.
REFERENCES
1. Pegg, A. E., and Williams-Ashman, H. G. (1981) in Polyamines in Biology and
Medicine (Morris, D. R., and Marton, L. J., eds), pp. 3–42, Marcel Dekker,
Inc., New York
2. Tabor, C. W., and Tabor, H. (1984) Annu. Rev. Biochem. 53, 749–790
3. Zappia, V., and Pegg, A. E. (1988) Progress in Polyamine Research. Novel
Biochemical Pharmacological and Clinical Aspects. Advances in Experi-
mental Medicine and Biology, Vol. 250, Plenum Press, New York
4. Pegg, A. E. (1986) Biochem. J. 234, 249–262
5. Metcalf, B. W., Bey, P., Danzin, C., Casera, P., and Vevert, J. P. (1978) J. Am.
Chem. Soc. 100, 2551–2553
6. Bacchi, C. J., and McCann, P. P. (1987) in Inhibition of Polyamine Metabolism
(McCann, P. P., Pegg, A. E., and Sjoerdsma, A., eds), pp. 317–344, Academic
Press, New York
7. Bitonti, A. J., McCann, P. P., and Sjoerdsma, A. (1987) Exp. Parasitol. 64,
237–243
8. Gillin, F. D., Reiner, D., and McCann, P. P. (1983) J. Protozool. 31, 161–163
9. Kaur, K., Emmett, K., McCann, P. P., Sjoerdsma, A., and Ullman, B. (1986)
J. Protozool. 33, 518–521
10. Giffin, B. F., McCann, P. P., Bitonti, A. J., and Bacchi, C. J. (1986) J. Protozool.
33, 238 –243
11. Cruz, A., and Beverley, S. M. (1990) Nature 348, 171–173
12. Tobin, J. F., Laban, A., and Wirth, D. F. (1991) Proc. Natl. Acad. Sci. U. S. A.
88, 864 868
13. Phillips, M. A., Coffino, P., and Wang, C. C. (1987) J. Biol. Chem. 262,
8721–8727
14. Hanson, S., Adelman, J., and Ullman, B. (1992) J. Biol. Chem. 267, 2350–2359
15. Wilson, K., Collart, F. R., Huberman, E., Stringer, J. R., and Ullman, B. (1991)
J. Biol. Chem. 266, 1665–1671
16. Allen, T., and Ullman, B. (1993) Nucleic Acids Res. 21, 5431–5438
17. Iovannisci, D. M., and Ullman, B. (1983) J. Parasitol. 69, 633–636
18. LeBowitz, J. H., Coburn, C. M., McMahon-Pratt, D., and Beverley, S. M. (1990)
Proc. Natl. Acad. Sci. U. S. A. 87, 9736 –9740
19. Cruz, A., Coburn, C. M., and Beverley, S. M. (1991) Proc. Natl. Acad. Sci.
U. S. A. 88, 7170–7174
20. Callahan, H. L., and Beverley, S. M. (1992) J. Biol. Chem. 267, 24165–24168
21. Sanger, F., Nicklen, S., and Coulson, A. R. (1977) Proc. Natl. Acad. Sci. U. S. A.
74, 5463–5467
22. Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular Cloning: A
Laboratory Manual, Cold Spring Harbor Laboratory, Cold Spring Harbor,
NY
23. Laemmli, U. K. (1970) Nature 227, 680685
24. Shim, H., and Fairlamb, A. H. (1988) J. Gen. Microbiol. 134, 807–817
25. Rojas-Chaves, M., Hellmund, C., and Walters, R. D. (1996) Mol. Biochem.
Parasitol. 75, 261–264
26. Fairlamb, A. H., Henderson, G. B., Bacchi, C. J., and Cerami, A. (1987) Mol.
Biochem. Parasitol. 24, 185–191
27. Fairlamb, A. H., Henderson, G. B., and Cerami, A. (1986) Mol. Biochem.
Parasitol. 21, 247–257
28. Henderson, G. B., Fairlamb, A. H., and Cerami, A. (1987) Mol. Biochem.
Parasitol. 24, 39–45
29. Kierszenbaum, F., Wirth, J. J., McCann, P. P., and Sjoerdsma, A. (1987) Proc.
Natl. Acad. Sci. U. S. A. 84, 42784282
30. Majumder, S., Wirth, J. J., Bitonti, A. J., McCann, P. P., and Kierszenbaum, F.
(1992) J. Parasitol. 78, 371–374
31. Hunter, K. J., Le Quesne, S. A., and Fairlamb, A. H. (1994) Eur. J. Biochem.
226, 1019 –1027
32. Ariyanayagam, M. R., and Fairlamb, A. H. (1997) Mol. Biochem. Parasitol. 84,
111–121
33. Li, F., Hua, S. B., Wang, C. C., and Gottesdiener, K. M. (1996) Mol. Biochem.
ODC Deficiency in L. donovani 3787
Parasitol. 78, 227–236
34. Steglich, C., and Scheffler, I. E. (1982) J. Biol. Chem. 257, 4603–4609
35. Paulus, T. J., Kiyona, P., and Davis, R. H. (1982) J. Bacteriol. 152, 291–297
36. Hafner, E. W., Tabor, C. W., and Tabor, H. (1979) J. Biol. Chem. 254,
12419–12426
37. Cohn, M. S., Tabor, C. W., and Tabor, H. (1980) J. Bacteriol. 142, 791–799
38. Casero, R. A., Jr., and Pegg, A. E. (1993) FASEB J. 7, 653–661
39. Wittich, R. M., and Walter, R. D. (1990) Mol. Biochem. Parasitol. 38, 13–18
40. Tabor, C. W. (1968) Biochem. Biophys. Res. Commun. 30, 339–342
41. Coward, J. K., and Pegg, A. E. (1987) Adv. Enzyme Regul. 26, 107–113
42. Titus, R. G., Gueiros-Filho, F. J., De Freitas, L. A. R., and Beverley, S. M.
(1995) Proc. Natl. Acad. Sci. U. S. A. 92, 10267–10271
ODC Deficiency in L. donovani3788
... The polyamine biosynthetic pathway of L. donovani consists of three enzymes: ODC, S-adenosyl methionine decarboxylase (AdoMetDC), and SpdS (Colotti and Ilari, 2011;Roberts et al., 2001). Extensive gene deletion studies resulted in the essentiality of this pathway in Leishmania ( Jiang et al., 1999;Reguera et al., 2009;Roberts et al., 2002, Roberts, 2004. There is a lack of spermine in Leishmania as the gene encoding spermine synthase is absent in the leishmanial genome (L.major Gene DB) ( Jiang et al., 1999), while spermine is a major polyamine in the mammalian hosts. ...
... Extensive gene deletion studies resulted in the essentiality of this pathway in Leishmania ( Jiang et al., 1999;Reguera et al., 2009;Roberts et al., 2002, Roberts, 2004. There is a lack of spermine in Leishmania as the gene encoding spermine synthase is absent in the leishmanial genome (L.major Gene DB) ( Jiang et al., 1999), while spermine is a major polyamine in the mammalian hosts. It has been known that putrescine is synthesized by the decarboxylation of ornithine in Leishmania and T. brucei; but in T. cruzi it totally depends on polyamine transport due to the absence of ODC. ...
... Moreover, the inhibition of polyamine transport showed potential trypanocidal effect (Reigada et al., 2018). The important roles of ODC (Boitz et al., 2009;Jiang et al., 1999), AdoMetDC (Roberts et al., 2002), or SpdS in parasite's viability and proliferation were established by generating the null mutants of these enzymes in L. donovani, and the significant differences between the polyamine pathways of the parasites and humans were explored by the phenotypic characterization of null mutants ( Jiang et al., 1999;Roberts et al., 2002). Therefore, these enzymes in Leishmania and other parasites can be the potential drug targets and have been validated using different enzyme inhibitors (Bacchi and Yarlett, 2002;Chaudhuri et al., 2006;Heby et al., 2003). ...
Article
Trypanosomatids are mainly responsible for heterogeneous parasitic diseases: Leishmaniasis, Sleeping sickness, and Chagas disease and control of these diseases implicates serious challenges due to the emergence of drug resistance. Redox-active biomolecules are the endogenous substances in organisms, which play important role in the regulation of redox homeostasis. The redox-active substances like glutathione, trypanothione, cysteine, cysteine persulfides, etc., and other inorganic intermediates (hydrogen peroxide, nitric oxide) are very useful as defence mechanism. In the present review, the suitability of trypanothione and other essential thiol molecules of trypanosomatids as drug targets are described in Leishmania and Trypanosoma. We have explored the role of tryparedoxin, tryparedoxin peroxidase, ascorbate peroxidase, superoxide dismutase, and glutaredoxins in the anti-oxidant mechanism and drug resistance. Up-regulation of some proteins in trypanothione metabolism helps the parasites in survival against drug pressure (sodium stibogluconate, Amphotericin B, etc.) and oxidative stress. These molecules accept electrons from the reduced trypanothione and donate their electrons to other proteins, and these proteins reduce toxic molecules, neutralize reactive oxygen, or nitrogen species; and help parasites to cope with oxidative stress. Thus, a better understanding of the role of these molecules in drug resistance and redox homeostasis will help to target metabolic pathway proteins to combat Leishmaniasis and trypanosomiases.
... The polyamine biosynthetic pathway of L. donovani consists of three enzymes: ODC, S-adenosyl methionine decarboxylase (AdoMetDC), and SpdS (Colotti and Ilari, 2011;Roberts et al., 2001). Extensive gene deletion studies resulted in the essentiality of this pathway in Leishmania ( Jiang et al., 1999;Reguera et al., 2009;Roberts et al., 2002, Roberts, 2004. There is a lack of spermine in Leishmania as the gene encoding spermine synthase is absent in the leishmanial genome (L.major Gene DB) ( Jiang et al., 1999), while spermine is a major polyamine in the mammalian hosts. ...
... Extensive gene deletion studies resulted in the essentiality of this pathway in Leishmania ( Jiang et al., 1999;Reguera et al., 2009;Roberts et al., 2002, Roberts, 2004. There is a lack of spermine in Leishmania as the gene encoding spermine synthase is absent in the leishmanial genome (L.major Gene DB) ( Jiang et al., 1999), while spermine is a major polyamine in the mammalian hosts. It has been known that putrescine is synthesized by the decarboxylation of ornithine in Leishmania and T. brucei; but in T. cruzi it totally depends on polyamine transport due to the absence of ODC. ...
... Moreover, the inhibition of polyamine transport showed potential trypanocidal effect (Reigada et al., 2018). The important roles of ODC (Boitz et al., 2009;Jiang et al., 1999), AdoMetDC , or SpdS in parasite's viability and proliferation were established by generating the null mutants of these enzymes in L. donovani, and the significant differences between the polyamine pathways of the parasites and humans were explored by the phenotypic characterization of null mutants ( Jiang et al., 1999;Roberts et al., 2002). Therefore, these enzymes in Leishmania and other parasites can be the potential drug targets and have been validated using different enzyme inhibitors (Bacchi and Yarlett, 2002;Chaudhuri et al., 2006;Heby et al., 2003). ...
Article
African protected areas strive to conserve the continent's great biodiversity with a targeted focus on the flagship 'Big Five' megafauna. Though often not considered, this biodiversity protection also extends to the lesser-known microbes and parasites that are maintained in these diverse ecosystems, often in a silent and endemically stable state. Climate and anthropogenic change, and associated diversity loss, however, are altering these dynamics leading to shifts in ecological interactions and pathogen spill over into new niches and hosts. As many African protected areas are bordered by game and livestock farms, as well as villages, they provide an ideal study system to assess infection dynamics at the human-livestock-wildlife interface. Here we review five zoonotic, multi-host diseases (bovine tuberculosis, brucellosis, Rift Valley fever, schistosomiasis and cryptosporidiosis)-the 'Microscopic Five'-and discuss the biotic and abiotic drivers of parasite transmission using the iconic Kruger National Park, South Africa, as a case study. We identify knowledge gaps regarding the impact of the 'Microscopic Five' on wildlife within parks and highlight the need for more empirical data, particularly for neglected (schistosomiasis) and newly emerging (cryptosporidiosis) diseases, as well as zoonotic disease risk from the rising bush meat trade and game farm industry. As protected areas strive to become further embedded in the socio-economic systems that surround them, providing benefits to local communities, One Health approaches can help maintain the ecological integrity of ecosystems, while protecting local communities and economies from the negative impacts of disease.
... For example, spermidine and spermine predominate in mammalian cells, while putrescine and spermidine are more abundant in trypanosomatids and other single cell organisms [56]. The polyamine spermine is neither produced nor utilized in trypanosomatids and the back-conversion of spermidine to putrescine does not exist [57][58][59]. Furthermore, while ODC and ADOMETDC are rapidly turned over in mammalian cells, these enzymes have a much longer half-life in trypanosomatid parasites [32,33,[60][61][62]. ...
... Deletion of the ODC gene in L. donovani led to growth arrest in promastigotes, axenic, and intracellular amastigotes, and it profoundly reduced infectivity in mice [57,107,111,127]. When ODC gene deletion mutants were incubated in polyamine-free media, intracellular putrescine pools were rapidly depleted while levels of spermidine, though initially decreased, were sustained at stable levels for over 12 days [57,127]. ...
... Deletion of the ODC gene in L. donovani led to growth arrest in promastigotes, axenic, and intracellular amastigotes, and it profoundly reduced infectivity in mice [57,107,111,127]. When ODC gene deletion mutants were incubated in polyamine-free media, intracellular putrescine pools were rapidly depleted while levels of spermidine, though initially decreased, were sustained at stable levels for over 12 days [57,127]. Although it remains unclear how spermidine levels were maintained in these mutants, it is possible that the back-conversion of trypanothione to spermidine via the bifunctional trypanothione synthetase-amidase helps to maintain intracellular spermidine levels. ...
Article
Full-text available
Parasites of the genus Leishmania cause a variety of devastating and often fatal diseases in humans and domestic animals worldwide. The need for new therapeutic strategies is urgent because no vaccine is available, and treatment options are limited due to a lack of specificity and the emergence of drug resistance. Polyamines are metabolites that play a central role in rapidly proliferating cells, and recent studies have highlighted their critical nature in Leishmania. Numerous studies using a variety of inhibitors as well as gene deletion mutants have elucidated the pathway and routes of transport, revealing unique aspects of polyamine metabolism in Leishmania parasites. These studies have also shed light on the significance of polyamines for parasite proliferation, infectivity, and host–parasite interactions. This comprehensive review article focuses on the main polyamine biosynthetic enzymes: ornithine decarboxylase, S-adenosylmethionine decarboxylase, and spermidine synthase, and it emphasizes recent discoveries that advance these enzymes as potential therapeutic targets against Leishmania parasites.
... Furthermore, it should be noted that ODC is not a target in T. cruzi, due to the absence of the gene encoding the enzyme in this parasite. In Leishmania, although ODC has been shown to be a pharmacotherapeutic target by genetic studies [162], the efficacy of DFMO and analogs is just observable in free-living forms, promastigotes [82], but this compound failed in intramacrophagic amastigotes and in vivo infections [163]. Other drug scaffolds to target ODC have been developed, some displaying ornithine (2) or putrescine (3) analogies [164,165] or owing to the structural dissimilarity between the ODC of humans and L. donovani (4), with different results [166]. ...
Article
Full-text available
Neglected tropical diseases transmitted by trypanosomatids include three major human scourges that globally affect the world’s poorest people: African trypanosomiasis or sleeping sickness, American trypanosomiasis or Chagas disease and different types of leishmaniasis. Different metabolic pathways have been targeted to find antitrypanosomatid drugs, including polyamine metabolism. Since their discovery, the naturally occurring polyamines, putrescine, spermidine and spermine, have been considered important metabolites involved in cell growth. With a complex metabolism involving biosynthesis, catabolism and interconversion, the synthesis of putrescine and spermidine was targeted by thousands of compounds in an effort to produce cell growth blockade in tumor and infectious processes with limited success. However, the discovery of eflornithine (DFMO) as a curative drug against sleeping sickness encouraged researchers to develop new molecules against these diseases. Polyamine synthesis inhibitors have also provided insight into the peculiarities of this pathway between the host and the parasite, and also among different trypanosomatid species, thus allowing the search for new specific chemical entities aimed to treat these diseases and leading to the investigation of target-based scaffolds. The main molecular targets include the enzymes involved in polyamine biosynthesis (ornithine decarboxylase, S-adenosylmethionine decarboxylase and spermidine synthase), enzymes participating in their uptake from the environment, and the enzymes involved in the redox balance of the parasite. In this review, we summarize the research behind polyamine-based treatments, the current trends, and the main challenges in this field.
... The polyamine biosynthesis pathway [13] is a promising route to drug development targeting arginase [14] and ornithine decarboxylase (ODC) [15]. We previously explored the synthesis of several scaffolds, such as cinnamic acid, pyrazolopyrimidines, and phenylhydrazides, using Leishmania arginase as a target. ...
Article
Full-text available
Leishmaniasis is a neglected tropical disease affecting millions of people worldwide. A centenary approach to antimonial-based drugs was first initiated with the synthesis of urea stibamine by Upendranath Brahmachari in 1922. The need for new drug development led to resistance toward antimoniates. New drug development to treat leishmaniasis is urgently needed. In this way, searching for new substances with antileishmanial activity, we synthesized ten anthranyl phenylhydrazide and three quinazolinone derivatives and evaluated them against promastigotes and the intracellular amastigotes of Leishmania amazonensis. Three compounds showed good activity against promastigotes 1b, 1d, and 1g, with IC50 between 1 and 5 μM. These new phenylhydrazides were tested against Leishmania arginase, but they all failed to inhibit this parasite enzyme, as we have shown in a previous study. To explain the possible mechanism of action, we proposed the enzyme PTR1 as a new target for these compounds based on in silico analysis. In conclusion, the new anthranyl hydrazide derivatives can be a promising scaffold for developing new substances against the protozoa parasite.
... It may be related to the difference in the Larginine metabolic pathway in this Leishmania species, where arginase does not seem essential for polyamine production. In studies with L. donovani, a related species that cause VL, arginase-deleted amastigotes survive within the cell without polyamine supplement but not promastigotes [83][84][85]. ...
Article
Full-text available
Leishmaniases are diseases caused by several Leishmania species, and many factors contribute to the development of the infection. Because the adaptive immune response does not fully explain the outcome of Leishmania infection and considering that the initial events are crucial in the establishment of the infection, we investigated one of the growth factors, the insulin-like growth factor-I (IGF-I), found in circulation and produced by different cells including macrophages and present in the skin where the parasite is inoculated. Here, we review the role of IGF-I in leishmaniasis experimental models and human patients. IGF-I induces the growth of different Leishmania species in vitro and alters the disease outcome increasing the parasite load and lesion size, especially in L. major- and L. amazonensis-infected mouse leishmaniasis. IGF-I affects the parasite interacting with the IGF-I receptor present on Leishmania. During Leishmania-macrophage interaction, IGF-I acts on the arginine metabolic pathway, resulting in polyamine production both in macrophages and Leishmania. IGF-I and cytokines interact with reciprocal influences on their expression. IL-4 is a hallmark of susceptibility to L. major in murine leishmaniasis, but we observed that IGF-I operates astoundingly as an effector element of the IL-4. Approaching human leishmaniasis, patients with mucosal, disseminated, and visceral diseases presented surprisingly low IGF-I serum levels, suggesting diverse effects than parasite growth. We observed that low IGF-I levels might contribute to the inflammatory response persistence and delayed lesion healing in human cutaneous leishmaniasis and the anemia development in visceral leishmaniasis. We must highlight the complexity of infection revealed depending on the Leishmania species and the parasite’s developmental stages. Because IGF-I exerts pleiotropic effects on the biology of interaction and disease pathogenesis, IGF-I turns up as an attractive tool to explore biological and pathogenic processes underlying infection development. IGF-I pleiotropic effects open further the possibility of approaching IGF-I as a therapeutical target.
... Isoprenoid and sterol synthesis Squalene synthase, sterol 14a-demethylase Kumar Saha et al. (1986) and Pomel et al. (2015) Folate metabolism Thymidylate synthase, dihydrofolate reductase and pteridine reductase Nare et al. (1997) Polyamine metabolism Hypusine pathway, arginase, and ornithine decarboxylase Jiang et al. (1999) Antioxidant metabolism and Detoxification ...
Article
The mortality rate of leishmaniasis is increasing at an alarming rate and is currently second to malaria amongst the other neglected tropical diseases. Unfortunately, many governments and key stakeholders are not investing enough in the development of new therapeutic interventions. The available treatment options targeting different pathways of the parasite have seen inefficiencies, drug resistance, and toxic side effects coupled with longer treatment durations. Numerous studies to understand the biochemistry of leishmaniasis and its pathogenesis have identified druggable targets including ornithine decarboxylase, trypanothione reductase, and pteridine reductase, which are relevant for the survival and growth of the parasites. Another plausible target is the sterol biosynthetic pathway; however, this has not been fully investigated. Sterol biosynthesis is essential for the survival of the Leishmania species because its inhibition could lead to the death of the parasites. This review seeks to evaluate how critical the enzymes involved in sterol biosynthetic pathway are to the survival of the leishmania parasite. The review also highlights both synthetic and natural product compounds with their IC50 values against selected enzymes. Finally, recent advancements in drug design strategies targeting the sterol biosynthesis pathway of Leishmania are discussed.
Article
The protozoan parasite Leishmania causes the tropical illness known as leishmaniases. Visceral Leishmaniasis (VL), alias kala-azar in Asian nations, one of the major diverse kinds, along with Cutaneous Leishmaniasis (CL) and Mucocutaneous Leishmaniasis (ML). Chemotherapy is the only choice used in the control of this ailment. The present leishmaniasis treatements have a number of side effects, a lengthy course of therapy, regional variation in efficacy, and the evolution of resistance; it is urgently necessary to find safer and more effective treatments. Finding a suitable pharmacological target along a biological pathway is the first stage in the drug discovery process. In this review, we will address some of the essential metabolic pathways that are possible pharmacological targets along with some important treatment options for leishmaniasis.
Chapter
Leishmaniasis is an infectious disease classified by WHO as one of the neglected tropical diseases. Due to the lack of human vaccines, chemotherapeutic agents represent the only strategy for disease combat. However, the current treatment is marked by variable efficacy, high toxicity, and high cost. Thus, the search for more efficient antileishmanial agents becomes urgent. Several studies carrying out the discovery or development of potent inhibitors of key enzymes of Leishmania metabolism have demonstrated promising results. The polyamine and trypanothione pathways are essential for parasite survival and pathogenesis. Polyamine synthesis allows parasite growth and influences infectivity. Moreover, the final product of the polyamine pathway spermidine is required for the synthesis of trypanothione, a scavenger of reactive oxygen and nitrogen species, which is essential for the maintenance of Leishmania redox balance. In the present chapter, the advances in the use of synthetic and natural inhibitors of the polyamine and trypanothione pathways from Leishmania are discussed.
Article
Background: There are three epidemiological types of visceral leishmaniasis in China, which are caused by Leishmania strains belonging to the L. donovani complex. The mechanisms underlying their differences in the population affected, disease latency, and animal host, etc., remain unclear. We investigated the protein abundance differences among Leishmania strains isolated from three types of visceral leishmaniasis endemic areas in China. Methods: Promastigotes of the three Leishmania strains were cultured to the log phase and harvested. The protein tryptic digests were analyzed with liquid chromatography-electrospray ionization-tandem mass spectrometry (LC-ESI-MS/MS), followed by label-free quantitative analysis. The MS experiment was performed on a Q Exactive mass spectrometer. Raw spectra were quantitatively analyzed with the MaxQuant software (ver 1.3.0.5) and matched with the reference database. Differentially expressed proteins were analyzed using the bioinformatics method. The MS analysis was repeated three times for each sample. Results: A total of 5012 proteins were identified across the KS-2, JIASHI-5 and SC6 strains in at least 2 of the three samples replicate. Of them, 1758 were identified to be differentially expressed at least between 2 strains, including 349 with known names. These differentially expressed proteins with known names are involved in biological functions such as energy and lipid metabolic process, nucleotide acid metabolic process, amino acid metabolic process, response to stress, cell membrane/cytoskeleton, cell cycle and proliferation, biological adhesion and proteolysis, localization and transport, regulation of the biological process, and signal transduction. Conclusion: The differentially expressed proteins and their related biological functions may shed light on the pathogenicity of Leishmania and targets for the development of vaccines and medicines.
Article
Full-text available
Strains of Escherichia coli K12 have been constructed which do not contain any of the polyamines normally present in a wild type strain, namely, 1,4-diaminobutane (putrescine) and spermidine. This phenotype arises as a consequence of the assembly into these strains of deletion mutations in speA (arginine decarboxylase), speB (agmatine ureohydrolase), speC (ornithine decarboxylase), and speD (adenosylmethionine decarboxylase). The polyamine-deficient strains grow indefinitely in the absence of polyamines but with a growth rate one-third of that found in the presence of polyamines. These strains can act as hosts for bacteriophages T4, T7, and f2, although the latter phage is poorly adsorbed; they can also maintain F' factors, ColE1 and P1 plasmids, and lysogeny by bacteriophage lambda. In contrast, the production of bacteriophage lambda in the absence of polyamines is strikingly decreased (greater than 99%) either after infection of a nonlysogen or after induction of a lysogen. A polyamine-deficient Hfr strain can transfer its chromosome to a recipient at a normal rate, but the number of recombinants observed in a cross is decreased approximately 300-fold. No such effect is observed when only the F- recipient strain in a cross is polyamine deficient.
Book
This book contains the scientific contributions presented at an International Symposium held in Sorrento, Italy, in June 1988 under the auspices of the University of Naples, the Italian Society of Biochemistry, and the National Research Council. The modern history of polyamines dates back to 1958 when the Tabors and Rosenthal first described the outlines of their biological synthesis. From then on, and particularly in the last ten years, a veri table explosion of Literature, characterized by thousands of papers, has witnessed the interest of the scientific community toward these molecules. Perhaps the old statement that "polyamines are molecules in search of a function" is no longer true today. A large number of effects exerted by these simple molecules are well known, and in many cases the mechanisms underlying these effects have been elucidated. The first section of the volume is entirely devoted to the enzymology and molecular biology of ornithine decarboxylase. Since its discovery by Gale more than forty years ago, this can be considered among the most widely studied enzymes in biology, and one of the most complex models in enzyme regulation. The mechanism of control of the enzyme activity at the transcriptional, post-transcriptional and post-translational levels, as well as the fine regulation by antizyme, are discussed in detail. The second group of contributions deals wi th AdoMet decarboxylase, propylamine transferase, polyamine oxidase and the other enzymes related to polyamine interconversion and regeneration.
Article
Radiolabelling studies using tritiated omithine, arginine and lysine, together with the relevant amino acid decarboxylase enzyme assays, indicate that the epimastigote stage of Trypanosoma cruzi is unable to synthesise significant amounts of putrescine and cadaverine de novo, compared to the amounts of these diamines scavenged from the growth medium. Radiolabelled putrescine is readily incorporated into spermidine, spennine and the trypanosomatid-specific polyamine-glutathione conjugate trypanothione (N1, N8-bis(glutathionyl)spermidine). Likewise, radiolabelled cadaverine is incorporated into the analogous polyamines aminopropylcadaverine, bis(aminopropyl)cadaverine and another major unidentified component. Subsequent studies showed this major component to be a novel polyamine-thiol conjugate whose structure was confirmed by chemical synthesis to be N1,N9-bis(glutathionyl)aminopropylcadaverine (homotrypanothione). Kinetic analyses using recombinant T. cruzi trypanothione reductase demonstrated that homotrypanothione disulphide is readily reduced by this enzyme with kinetic parameters similar to trypanothione disulphide, suggesting that it is a physiological substrate in vivo. Thus the epimastigote form of T. cruzi differs significantly from the African trypanosomes and Leishmania in (a) being unable to synthesise significant amounts of diamines de novo, (b) converting significant amounts of putrescine and cadaverine to spermine and bis(aminopropyl)cadaverine, respectively and (c) the ability to synthesise homotrypanothione as well as trypanothione. The implications of these findings with respect to the prospective chemotherapy of Chagas’ disease are discussed.
Article
A trypanothione-dependent peroxidase activity has been identified in the insect trypanosomatid Crithidia fasciculata and in the mammalian trypanosome Trypanosoma brucei. Using organic hydroperoxides as oxidant, specific peroxidase activities in these organisms are 5.0 and 1.0 nmol min-1 (10(8) cells)-1 respectively. The T. brucei peroxidase had an activity of 0.4 nmol min-1 (10(8) cells)-1 using hydrogen peroxide as oxidant. The enzymeis specific for the N1,N8-bis(glutathionyl)spermidine conjugate (dihydro-trypanothione); N1-mono-glutathionylspermidine is not a substrate. Experiments to demonstrate that this parasite peroxidase may contain selenium were inconclusive. However, bloodstream T. brucei can incorporate radiolabelled selenite into proteins.