ArticlePDF Available

Complete development and long-term maintenance of Cryptosporidium parvum human and cattle genotypes in cell culture

Authors:

Abstract and Figures

This study describes the complete development (from sporozoites to sporulated oocysts) of Cryptosporidium parvum (human and cattle genotypes) in the HCT-8 cell line. Furthermore, for the first time the complete life cycle was perpetuated in vitro for up to 25 days by subculturing. The long-term maintenance of the developmental cycle of the parasite in vitro appeared to be due to the initiation of the auto-reinfection cycle of C. parvum. This auto-reinfection is characterised by the production and excystation of new invasive sporozoites from thin-walled oocysts, with subsequent maintenance of the complete life cycle in vitro. In addition, thin-walled oocysts of the cattle genotype were infective for ARC/Swiss mice but similar oocysts of the human genotype were not. This culture system will provide a model for propagation of the complete life cycle of C. parvum in vitro.
Content may be subject to copyright.
Complete development and long-term maintenance of Cryptosporidium
parvum human and cattle genotypes in cell culture
N.S. Hijjawi
a
, B.P. Meloni
b
, U.M. Morgan
a
, R.C.A. Thompson
a,
*
a
Division of Veterinary and Biomedical Sciences, Murdoch University, South Street, Murdoch, W.A. 6150, Australia
b
Sir Charles Gairdner Hospital, QEII Medical Centre, Nedlands, W.A. 6009, Australia
Received 1 February 2001; received in revised form 15 March 2001; accepted 15 March 2001
Abstract
This study describes the complete development (from sporozoites to sporulated oocysts) of Cryptosporidium parvum (human and cattle
genotypes) in the HCT-8 cell line. Furthermore, for the ®rst time the complete life cycle was perpetuated in vitro for up to 25 days by
subculturing. The long-term maintenance of the developmental cycle of the parasite in vitro appeared to be due to the initiation of the auto-
reinfection cycle of C.parvum. This auto-reinfection is characterised by the production and excystation of new invasive sporozoites from
thin-walled oocysts, with subsequent maintenance of the complete life cycle in vitro. In addition, thin-walled oocysts of the cattle genotype
were infective for ARC/Swiss mice but similar oocysts of the human genotype were not. This culture system will provide a model for
propagation of the complete life cycle of C.parvum in vitro. q2001 Australian Society for Parasitology Inc. Published by Elsevier Science
Ltd. All rights reserved.
Keywords:Cryptosporidium parvum; Complete life cycle; In vitro development; Human genotype; Cattle genotype; Auto-reinfection; Thin-walled oocyst;
HCT-8 cell line
1. Introduction
Cryptosporidium parvum is an enteric parasite that causes
diarrhoeal disease in humans and domesticated animals
world wide (O'Donoghue, 1995). The parasite infects intest-
inal epithelial cells resulting in self-limiting diarrhoea in
immunocompetent persons. However, it is more severe,
and potentially fatal, to the immunosuppressed population,
especially those with acquired immunode®ciency syndrome
(AIDS) (Peterson, 1992; O'Donoghue, 1995). Although a
large number of anti-parasitic drugs have been tested against
Cryptosporidium, no consistently effective chemotherapeu-
tic agent is available and a healthy, intact immune system
remains the only reliable defence (Theodos et al., 1998;
Tzipori, 1998).
There have been more than 25 reports describing cell
lines that support the growth of both sexual and asexual
stages of C.parvum in vitro. These studies showed that
different cells can be infected with C.parvum, in which
maximum growth was reached after 48±72 h, but then
gradually declined (Eggleston et al., 1994; Lawton et al.,
1997; Tzipori, 1998). Apart from the production of small
numbers of oocysts in some cell lines, it has not been possi-
ble to initiate re-infection of the cell line from these oocysts
(Lawton et al., 1997). The failure to perpetuate the infection
in vitro has been attributed to the lack of thin-walled oocyst
production, which is thought to be essential for the auto-
infection cycle in vivo (Current and Garcia, 1991).
Recent studies have shown that C.parvum is not a geneti-
cally uniform species but encompasses at least seven
distinct genotypes that appear to be host speci®c; human,
cattle, pig, marsupial, dog, ferret and mouse (Awad-el-
Kariem et al., 1995; Bonnin et al., 1996; Peng et al.,
1997; Morgan et al., 1995, 1997, 1998, 1999; Xiao et al.,
1999; Sulaiman et al., 2000). Humans are susceptible to
infection with the human and the cattle genotypes of C.
parvum, providing evidence of zoonotic transmission.
Regarding in vitro culturing of C.parvum, most success
has been obtained using cattle isolates (Villacorta et al.,
1996; Yang et al., 1996; Deng and Cliver, 1998), although
two attempts using oocysts from human patients with AIDS
have also been reported (Current and Haynes, 1984; Burand
et al., 1991). However, with the exception of the study by
Meloni and Thompson (1996) all C.parvum cultures have
been established without prior genotyping of the isolate
used.
International Journal for Parasitology 31 (2001) 1048±1055
0020-7519/01/$20.00 q2001 Australian Society for Parasitology Inc. Published by Elsevier Science Ltd. All rights reserved.
PII: S0020- 7519(01)00212-0
www.parasitology-online.com
* Corresponding author. Fax: 161-8-9360-2466.
E-mail address: andrew_t@numbat.murdoch.edu.au
(R.C.A. Thompson).
The establishment of the complete life cycle of C.parvum
in vitro will allow the evaluation of drug therapies against
different life-cycle stages, enhance research into the biology
of host cell±parasite interactions and enable ampli®cation of
parasite material for further immunological, biochemical
and molecular studies. It could also be used for the assess-
ment of viability of C.parvum oocysts isolated from envir-
onmental samples (Deng and Cliver, 1998; Di Giovanni et
al., 1999; Rochelle et al., 1999). Finally, the value of an in
vitro culture system for C.parvum will be further enhanced
if complete development of both human infective genotypes
(human and cattle) can be propagated.
2. Materials and methods
2.1. Parasite isolates
The C.parvum cattle isolate (Swiss cattle C26) used
during this study was originally obtained from the Institute
of Parasitology, Zurich and has subsequently been passaged
through mice as previously described by Meloni and
Thompson (1996). The human isolate was obtained from a
local diagnostic laboratory from a patient in Perth, Western
Australia.
2.2. Genotyping of parasite material
The human and cattle genotypes of C.parvum were iden-
ti®ed according to the method described by Morgan et al.
(1997) based on direct PCR analysis and sequencing of the
18S rDNA. Genotyping of the human isolate was carried out
using DNA extracted from faecal oocysts and from parasite
material collected from HCT-8-infected cultures. The cattle
isolate (Swiss cattle C26) was previously genotyped by
Morgan et al. (1997) and furthermore, genotyping was
carried out for oocysts puri®ed from mice after being
infected from in vitro-derived oocysts of cattle origin.
2.2.1. Direct PCR analyses
Speci®c PCR primers that directly differentiate between
human and cattle genotypes on the basis of the size of the
PCR product were also used in the present study (Morgan et
al., 1997). The human-speci®c primer, CP-HR, which
ampli®es a 411-bp product from human isolates only, and
the cattle-speci®c primer, CP-CR, which ampli®es a 312-bp
product from animal isolates only, were used for the PCR
ampli®cation. The ampli®cation products were subjected to
electrophoretic separation in 1.5% agarose gels, stained
with ethidium bromide, visualised under UV light and
compared with positive controls of both cattle and human
genotypes.
2.2.2. Sequencing of the 18S rDNA
For further con®rmation of the human C.parvum geno-
type, DNA extracted from oocysts and infected HCT-8
cultures was used to amplify the 18S rDNA for sequencing
(Morgan et al., 1997).
2.3. Puri®cation of C. parvum oocysts
For in vitro culturing, C.parvum oocysts of the human
genotype were puri®ed from faeces using a routine puri®ca-
tion procedure (Morgan et al., 1995). Cryptosporidium
parvum oocysts of the cattle genotype were obtained by
infecting 7±8-day-old ARC/Swiss mice with 100 000±
120 000 oocysts. The oocysts were isolated and puri®ed
according to the procedure described by Meloni and
Thompson (1996), with the addition of a ®nal bleaching
step. On day 8 p.i., mice were euthanised (CO
2
inhalation)
and the jejunum, ileum, caecum, colon and rectum removed,
placed in sterile PBS/0.02% Tween-20 <4 ml/mouse) and
dissected into small segments. The segments were further
homogenised at 48C and sputasol (0.005 g/ml of the suspen-
sion) was added. The homogenate was then left at room
temperature (RT) for 90±120 min on a rotary mixer before
centrifugation at 2000 £gfor 8 min. The supernatant was
removed, 40 ml of PBS/0.02% Tween-20 and 10 ml of ether
added before mixing vigorously (20±30 s) and centrifuging
for 8 min at 2000 £g. The supernatant was removed and the
pellet resuspended in 4 ml PBS. For further puri®cation of
oocysts, the 4-ml suspension was layered on to a Ficoll
gradient (1%/0.5% Ficoll prepared in PBS containing 16%
sodium diatrizoate). Gradients were centrifuged at 2000 £g
for 30 min at RT. Oocysts were collected from the PBS/
0.5% Ficoll interface and washed twice with PBS and made
up to 10 ml with sterile PBS. They were then bleach treated
by adding sodium hypochlorite (200±300 ml/ml), and incu-
bated at RT for 20 min, washed twice with PBS and centri-
fuged at 2000 £gfor 8 min. Finally, puri®ed oocysts were
resuspended in cold sterile PBS and stored at 48C after
adding 15 ml/ml antibiotic solution containing ampicillin
(10 mg/ml) and lincomycin (4 mg/ml).
2.4. Pre-treatment of oocysts and culture media preparation
Oocysts were excysted to release sporozoites in a freshly
prepared, ®lter-sterilised (0.22 mm ®lter) excystation
medium composed of acidic H
2
O (pH 2.5±3) containing
0.5% trypsin and incubated in a water bath at 378C for 20
min with mixing every 5 min. Thereafter, the excystation
suspension was centrifuged at 2000 £gfor 4 min at RT.
Oocysts were resuspended in maintenance medium (100
ml RPMI-1640) containing 0.03 g l-glutamine, 0.3 g
sodium bicarbonate, 0.02 g bovine bile, 0.1 g glucose, 25
mg folic acid, 100 mg 4-aminobenzoic acid, 50 mg calcium
pantothenate, 875 mg ascorbic acid, 1% FCS, 15 mM
HEPES buffer, 10 000 units penicillin G and 0.01 g strep-
tomycin, adjusted to pH 7.4. The percent excystation, for the
human and cattle genotypes, was calculated by scoring at
least 300 oocysts as empty or intact after 3 h incubation in
maintenance medium. The percent excystation was calcu-
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±1055 1049
lated as the number of empty oocysts/number of empty 1
intact oocysts.
2.5. Preparation and infection of host cells
HCT-8 cells (ATCC; CCL244) were grown in RPMI-
1640, 10% FCS in 25-cm
2
¯asks and seeded 24 h prior to
infection to allow them to reach monolayer. HCT-8 cells
were infected by removing the existing media and adding
maintenance medium containing 50 000 pre-treated oocysts
(2000 oocysts/cm
2
). Flasks were kept at 378C in a candle jar
(18.8% O
2
, 1.97% CO
2
). During in vitro cultivation, the
medium was changed every 2±3 days to maintain the pH
within the range 7.2±7.6.
2.6. Subculturing to new cell monolayers
Subculturing was carried out by collecting supernatant or
scraped cells from infected cultures which were used to
inoculate a new cell monolayer as described above. After
2 h, the supernatant and the cell debris were removed by
washing the cell monolayer with pre-warmed PBS at pH 7.2,
resupplied with maintenance medium and the ¯asks incu-
bated at 378C in a candle jar where the media was changed
every 2±3 days. Supernatant only was used later in the study
since the cells and cell debris appeared to adhere ®rmly to
the cells in the new monolayer and interfere with parasite
development.
2.7. Puri®cation of thin-walled oocysts from in vitro
cultures
An attempt to purify the thin-walled oocysts from in vitro
cultures was carried out during the present study. Super-
natant was collected from culture ¯asks after passaging
the parasite for 12 days, centrifuged, pellets resuspended
in PBS (pH 7.2) and puri®ed using Ficoll gradient centrifu-
gation as described above.
2.8. Infectivity of culture-derived oocysts to mice
Culture medium from four 75-cm
2
culture ¯asks
(approximately 200 ml) was collected from 5-day-old
cultures infected with oocysts (2000 oocysts/cm
2
)ofC.
parvum (cattle and human genotypes). As a control,
RPMI-1640 maintenance medium containing 100 000 pre-
treated oocysts, were also incubated for 5 days under the
same conditions in cell-free ¯asks without media changes.
This was done to be sure that the infectivity in mice resulted
from the thin-walled oocysts produced in vitro and not from
the non-excysted oocyst which remained in culture from the
initial inoculum. The culture media from the infected mono-
layers and control ¯asks were centrifuged at 2000 £gfor 5
min and the pellet reconstituted in 2 ml PBS before being
inoculated intragastrically into 7±8-day-old ARC/Swiss
mice (100 ml/mouse). Eight days p.i. mice were processed
for oocyst puri®cation as described above.
2.9. Examination of HCT-8 cells infected with C. parvum
oocysts (human and cattle genotypes)
Cultures were examined daily at magni®cations ranging
from £150 to £600 using an inverted light microscope
(Olympus IMT-2) ®tted with a heating chamber. Nomarski
phase-contrast microscopy (Olympus BX50) and Optimas
image analysis (MS-DOS operating system) were used for
capturing images of C.parvum stages. Because of the dif®-
culty in capturing the images of the different stages, mono-
layers infected with C.parvum (cultured on 25-mm
2
coverslips or scraped intact from ¯asks) were compressed
onto a glass slide, examined and photographed under oil
immersion £1000 magni®cation).
3. Results
3.1. Genotyping
Analysis of the human isolate with the diagnostic primer
(CP-HR) ampli®ed a 411-bp product from faecal oocysts
and infected HCT-8 cultures after 72 h p.i. Furthermore,
sequencing of the 18S rDNA (DNA extracted from oocysts
and HCT-8 cultures after 72 h of infection) con®rmed that
this isolate exhibited the human genotype recognition
sequence according to Morgan et al. (1997). The genotype
of the cattle isolate, used during this study, which has been
continuously passaged in mice in our laboratory, was initi-
ally determined by Morgan et al. (1997) using the above two
methods and con®rmed regularly ever since.
3.2. Excystation of C. parvum oocysts prior to culturing
The excystation rate of the cattle and human genotypes
was 88 and 86%, respectively, after 3 h incubation in excys-
tation medium. Fig. 1a±d shows the sequential stages of
excystation of sporozoites from pre-treated oocysts of the
cattle genotype, which did not differ from the oocysts of the
human genotype.
3.3. Observation of HCT-8 cells infected with C. parvum
Human and cattle genotypes of C.parvum were observed
to complete their life cycle in HCT-8 cells (from sporozoites
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±10551050
Fig. 1. Nomarski interference-contrast photomicrographs of the steps of
excystation of Cryptosporidium parvum oocysts (cattle genotype) after
treatment with trypsin and bile salts. (a) Intact thick-walled oocyst before
the treatment. (b) Partially excysted oocyst with the suture(s) partially
opened. (c) Free sporozoite released from the suture. (d) Shell of an
empty oocyst with residuum (r) and an open suture(s). Bar: 5 mm.
to sporulated oocysts) under the culturing conditions
described above (Figs. 1±5). The various developmental
stages were maintained in culture for up to 25 days by the
passage of parasite life-cycle stages from the supernatant of
5-day-old infected monolayers to fresh HCT-8 cell mono-
layers.
3.4. Description of the various developmental stages of C.
parvum (human and cattle genotypes) in culture
During the ®rst 24 h p.i., intracellular circular forms
likely to be trophozoites/uninucleate meronts (Figs. 2 and
3) formed as an early stage in the life cycle of C.parvum.At
this time the infection appeared to be restricted to the site of
initial sporozoite infections (Fig. 2). From 48±72 h p.i.
approximately 70±80% of cells became infected and all
the developmental stages could be clearly identi®ed.
These included merozoites, meront I, meront II, macroga-
metes, microgametocytes and oocysts. No major differences
could be seen in the in vitro development of the life-cycle
stages of C.parvum between the human and cattle geno-
types. However, the time needed to complete the life cycle
was much shorter in the human genotype. Unsporulated and
sporulated oocysts were seen in situ after 72 h p.i. for the
human genotype and after 5 days p.i. for the cattle genotype.
The stages observed in culture were characterised as
follows.
3.4.1. Merozoites
The merozoites were usually identi®ed during their
attempts to penetrate cells after 48-h p.i. Merozoites are
thread-shaped with rounded anterior and posterior ends.
We consistently observed merozoites inside HCT-8 cells
(Fig. 4a,b) and while they were released from the cells.
Once merozoites were released they penetrated other cells
rapidly. The merozoites displayed vigorous gliding and ¯ex-
ing movements and continuously attempted to invade cells.
3.4.2. Trophozoites
Round or oval intracellular forms, 2:7£2:7mmin
diameter, and considered as a transitional stage from spor-
ozoites and merozoites to meronts (Figs. 2 and 3).
3.4.3. Meront I
Developing meronts, 3:75 £4mm in diameter, with six or
eight merozoites (Figs. 3 and 4a,b) were observed in culture
and on one occasion six merozoites were counted while
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±1055 1051
Fig. 2. Light photomicrograph of HCT-8 cell monolayer after 24 h of
infection with Cryptosporidium parvum (cattle genotype). Circular struc-
tures are trophozoite stage (t). Bar: 200 mm.
Fig. 3. Light photomicrograph of HCT-8 cells after 48 h infection with
Cryptosporidium parvum (cattle genotype). Note the presence of uninucle-
ate meront/trophozoite stage (t) and immature type I meront (m). Bar: 200
mm.
Fig. 4. Nomarski interference-contrast photomicrographs of stages of endo-
genous development of Cryptosporidium parvum in HCT-8 cell (both from
human and cattle genotypes). (a) Immature type I meront with eight mero-
zoites (m) in focus. (b) Mature type I meront with six merozoites (m). (c)
Immature type II meront with four of the nuclei in focus. (d) Microgamonts
with microgametocytes and a large residuum (r). (e) Macrogamont contains
a large eccentric nucleus (n). Bar: 5 mm.
excysting. The merozoites in type I meronts are arranged
parallel to each other like segments of an orange.
3.4.4. Meront II
Meront II, 3:1£2:8mm in diameter, can be differentiated
from meront I by having four merozoites but we never
observed the release of four merozoites from type II meronts
(Fig. 4c).
3.4.5. Microgamonts
Microgamonts, 5:6£3:4mm in diameter, were observed
containing 14±16 non-¯agellated microgametes (Fig. 4d)
occupying most of the cell around a residuum. Free micro-
gametes as described by Current and Reese (1986) were
bullet-shaped, and displayed a jerky gliding movement in
the culture supernatants presumably after disruption of the
microgamont. Budding on the surface of microgamonts was
observed, and this might be an attempt by microgametes to
escape from the host cell.
3.4.6. Macrogamonts
This stage was distinguished by being large in size 4£4
mm in diameter) and having a large peripheral nucleus (Fig.
4e).
3.4.7. Oocysts
Culture-derived oocysts lacked an outer thick wall, which
is usually present in oocysts puri®ed from faeces (Fig. 5g),
and were identi®ed as thin-walled oocysts with a 5 £5mm
diameter (Fig. 5d±f). Unsporulated, as well as sporulated
oocysts were observed in situ (Fig. 5a±c) and in culture
supernatant. Some oocysts were seen empty and some had
only two sporozoites and a residuum; however, some spor-
ozoites were observed while excysting from the thin-walled
oocysts (Fig. 5f). Thin-walled oocysts were puri®ed from
the culture supernatant using Ficoll gradient.
3.4.8. Sporozoites
Sporozoites were ®rst observed within thin-walled
oocysts 72 h p.i. and thereafter (Fig. 5b±f). They are 5:2£
1:2mm in diameter and characterised by having a comma
shape with a rounded posterior end and a pointed, tapered
anterior end. Sporozoites exhibited a gliding movement and
many were seen actively moving inside the oocyst and
excysting, and ¯oating in the culture media.
Measurements of stages observed in culture were made
on captured images and were within the ranges of these
stages described by Current and Reese (1986) in vivo.
3.5. Infectivity of thin-walled oocysts to mice
Culture-derived oocysts of the cattle genotype were
infective to 7±8-day-old ARC/Swiss mice. An oocyst
yield of 8:3£105(collected from six mice) was obtained
after puri®cation as described in Section 2. However, under
the same conditions, culture-derived oocysts of the human
genotype and from control ¯asks (cattle genotype oocysts
incubated without the HCT-8 cells) failed to infect mice.
Oocysts puri®ed from human faeces (human genotype) also
failed to infect mice.
4. Discussion
The present study describes, for the ®rst time, the complete
development and long-term maintenance of C.parvum of
both cattle and human genotypes in vitro. The human geno-
type, which is restricted to humans, has never been propa-
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±10551052
Fig. 5. Nomarski interference-contrast photomicrographs showing the stages of oocysts of Cryptosporidium parvum (cattle and human genotypes) develop-
ment in HCT-8 cell line. The earliest of these images was taken after 72 h p.i. (a) Unsporulated oocyst in intracellular location (in situ) with residuum (r). (b,c)
Sporulated oocyst in situ detected in HCT-8 cells after 72 h of infection with C.parvum (human genotype); note the presence of three sporozoites (sp) in focus
in (c). (d) Thin-walled oocyst, puri®ed using Ficoll gradient, from a 12-day-old culture (cattle genotype). (e) Fully sporulated thin-walled oocyst with four
sporozoites (sp) and residuum. (f) Sporozoite in an attempt to excyst from a thin-walled oocyst. (g) Thick-walled oocyst after puri®cation from faeces of a
human patient infected with C.parvum; note the presence of the outer thick-wall (ow) which is absent from the thin-walled oocyst in (d). Bar: 5 mm.
gated or maintained in culture for such a long period (25
days). Furthermore, apart from the study by Meloni and
Thompson (1996), no previous workers genotyped Cryptos-
poridium isolates prior to in vitro culturing. Although there
have been two studies describing the complete development
of C.parvum originating from human patients with AIDS
(Current and Haynes, 1984; Burand et al., 1991), neither of
these isolates were genotyped. A number of human cases
have been reported to be due to infection with the cattle
genotype (Morgan et al., 1997; Peng et al., 1997; Widmer
et al., 1998). Research in our laboratory, screening a large
number of faecal samples, has shown that approximately
17% of isolates of Cryptosporidium infecting humans
display the cattle genotype (Morgan et al., 1998). Further-
more, recent genetic evidence showed that patients with
AIDS might be susceptible to a wide range of Cryptospor-
idium species and genotypes to which an immunocompetent
individual is not susceptible to (Widmer et al., 1998; Morgan
et al., 2000). Consequently, genotyping should be an essen-
tial step before culturing different isolates of Cryptospori-
dium to allow interpretation of differences in their
development, pathogenesis, host cell interactions and
susceptibility to chemotherapeutic agents. Several attempts
to propagate the human genotype of C.parvum in neonatal
mice or cattle proved to be unsuccessful (Meloni and Thomp-
son, 1996; Peng et al., 1997; Widmer et al., 1998; present
study), although a recent study described the successful
propagation of the human genotype in gnotobiotic piglets
(Widmer et al., 2000). However, the maintenance of the
parasite in animal hosts is labour-intensive, expensive, and
it is not easy to purify the oocysts from large animals such as
pigs and calves. In addition, they will not support the growth
of the parasite for prolonged periods of time. Thus the present
in vitro culturing system should provide a model for the
maintenance and propagation of both the C.parvum human
and cattle genotypes.
In the present study, the in vitro cultivation of C.parvum
was based on the method of Meloni and Thompson (1996),
but with some modi®cations to pH. The pH appeared to play
an important role in successfully maintaining the growth of
C.parvum in vitro. The pH was monitored and maintained
within an optimum range (7.2±7.6). This was achieved by
changing the media regularly, every 2±3 days, and adding
HEPES buffer at 15 mM. An optimum pH is likely to be
important for both cell invasion and release of parasite stages
(Upton et al., 1995; Meloni and Thompson, 1996) as well as
for the stability of the host cell monolayer. Sloughing of the
host cell monolayer was observed on several occasions
following sudden drops or rises in pH. The low number of
progeny oocysts in vitro in previous studies might be due, at
least in part, to pH ¯uctuations. It has been suggested that
microgamete inactivity at suboptimal pH might be the reason
for the low yield of oocysts of C.parvum in vitro (Upton,
1997). With Plasmodium spp., it has been shown that micro-
gamete emergence and ex¯agellation is a pH dependent
process (Kamamoto et al., 1991). Furthermore, the yield of
Eimeria tenella oocysts in cell culture was increased three-
fold with better pH regulation (Doran, 1971).
A healthy, 1±4-day-old intact host cell monolayer also
appeared to be an important factor for C.parvum growth
in vitro. During the ®rst 4 days p.i., the infection reached
maximum levels, after which time large quantities of cell
debris were evident in the culture medium. The presence of
cell debris and the reduced ability of the merozoites/spor-
ozoites to infect old cells are likely factors contributing to
reduced parasite development. Older cells are known not to
be very susceptible to infection with species of coccidia
(Dvorak and Crane, 1981; Dvorak and Howe, 1977). During
the present study, the problem of overgrowth and degenera-
tion of host cells was avoided by subculturing. Multiple
subculturing proved to be successful and resulted in the
maintenance of the parasite in vitro for up to 25 days.
Previous failures to maintain the intensity of infection
with C.parvum in vitro have been attributed to the poor/
limited production of thin-walled oocysts which are consid-
ered to be an essential factor for auto-reinfection (Current
and Garcia, 1991). Earlier attempts at subculturing in our
laboratory using the cattle genotype were unsuccessful
(unpublished observations). This may have been attributed
to the time subculturing was carried out. Earlier attempts at
subculturing were performed at a time when oocysts were
unlikely to be present (2±3 days p.i.). These ®ndings suggest
that thin-walled oocysts, from which new invasive sporo-
zoites were released, and not merozoites are essential for
perpetuating the life cycle of C.parvum in vitro. Auto-rein-
fection through the production of thin-walled oocysts may
explain why Cryptosporidium infections in immunode®-
cient individuals develop into persistent, life-threatening
diarrhoea with infection of extra-intestinal sites such as
respiratory and biliary ducts (Current and Haynes, 1984).
Apart from Current and Haynes (1984), who described the
presence of some thick-walled oocysts in an intracellular
location in a human foetal lung cell line, in all other studies
including the present, C.parvum oocysts produced in vitro
were thin-walled (Fig. 5a±f). From the present study it
appears likely that such thin-walled oocysts are the stage
responsible for perpetuating the cycle in vivo. Failure in
the production of thick-walled oocysts in culture might be
due to the absence of an essential ingredient from the culture
medium (such as a speci®c enzyme, hormone or growth
factor, which is normally present in vivo) or the lack of an
effective immune response. It is possible that in vivo, the
parasite responds to the development of an immune response
by producing thick-walled oocysts, thus curtailing the auto-
infective cycle, which continues in individuals with impaired
immunity. This hypothesis could be tested by adding
immune serum or other immunological factors against C.
parvum to the culture medium. Many molecules on the
surface of sporozoites and merozoites are strongly immuno-
genic; however, the parasite appears to respond to this chal-
lenge (immune response) by developing rapidly into thick-
walled oocysts, which can infect new susceptible hosts.
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±1055 1053
The thin-walled oocysts (cattle genotype) derived from
culture were infective to mice and produced a yield of 8:3£
105oocysts (from six mice at 8 days p.i.), which is higher than
the initial inoculum (6 £105oocysts) used to infect the cell
monolayers. Culture-derived oocysts of the human genotype
failed to infect mice, a result consistent with earlier in vivo
studies which showed that the human genotype does not read-
ily infect mice (Current and Reese, 1986; Pozio et al., 1992;
Meloni and Thompson, 1996; Peng et al., 1997; Widmer et
al., 1998). Although in vivo infections with thick-walled
oocysts usually give a much higher yield (3±4 million
oocysts/mouse) (Meloni and Thompson, 1996), the future
standardisation of the in vitro technique described here
may lead to large-scale production of oocysts and obviate
the need to maintain the parasite in animal models.
During the present study, although no attempts were
made for the quantitation of the different developmental
stages encountered in vitro, the human genotype appeared
to be much more aggressive in its growth in HCT-8 cells
than the cattle genotype. After 72 h p.i. with oocysts of the
human genotype, HCT-8 cells appeared to be perforated
with large numbers of parasitic stages, especially mero-
zoites that could be seen continuously attempting to pene-
trate the cells. Furthermore, the human genotype completed
the life cycle with the production of oocysts within 72 h,
whereas oocysts of the cattle genotype could not be detected
in culture before 5 days p.i. The reason for this difference
might be that the human genotype has a faster growth rate
than the cattle genotype and/or is better adapted to grow in
the HCT-8 cell line, which is of human origin.
The culture system described in the present study for the
in vitro maintenance of C.parvum will provide a model to
investigate host cell±parasite interactions and drug ef®cacy
with different genotypes and life-cycle stages. It may also
prove valuable for investigating the pathogenesis of the
parasite in the absence of an effective immune response.
The system can also be used for the large-scale production
of oocysts and to assess the viability of C.parvum oocysts
from environmental samples.
Acknowledgements
We should like to thank Russ Hobbs for his excellent
assistance with the illustrations and Aileen Eliott for her
diagnostic expertise. We are also grateful to George Di
Giovanni for helpful discussions at the commencement of
these studies. The ®nancial assistance of Murdoch Univer-
sity and the Pig Research and Development Corporation is
acknowledged.
References
Awad-el-Kariem, F.M., Robinson, H.A., Dyson, D.A., Evans, D., Wright,
S., Fox, M.T., McDonald, V.M., 1995. Differentiation between human
and animal strains of Cryptosporidium parvum using isoenzyme typing.
Parasitology 110, 129±32.
Bonnin, A.M., Fourmaux, N., Dubremetz, J.F., Nelson, R.G., Gobet, P.,
Harly, G., Buisson, M., Puygauthier-Toubas, D., Gabriel-Pospisil, F.,
Naciri, M., Camerlynck, P., 1996. Genotyping human and bovine
isolates of Cryptosporidium parvum by polymerase chain reaction-
restriction fragment length polymorphism analysis of a repetitive
DNA sequence. FEMS Microbiol Lett. 137, 207±11.
Burand, M.F., Favennec, E., Bizet, L., Gobert, J., Deluol, A., 1991. Sexual
stage development of Cryptosporidium in the Caco-2 cell line. Infect.
Immunol. 59, 4610±3.
Current, L., Garcia, L.S., 1991. Cryptosporidiosis. Clin. Microbiol. Rev. 4,
325±58.
Current, W.L., Haynes, T.B., 1984. Complete development of Cryptospor-
idium in cell culture. Science 224, 603±5.
Current, W.L., Reese, N.C., 1986. A comparison of endogenous develop-
ment of three isolates of Cryptosporidium in suckling mice. J. Proto-
zool. 33, 98±108.
Deng, M.Q., Cliver, D.O., 1998. Cryptosporidium parvum development in
the BS-C-1 cell line. J. Parasitol. 84, 8±15.
Di Giovanni, G.D., Hashemi, F.H., Shaw, N.J., Abrams, F.A., Lechevallier,
M.W., Abbaszadegan, M., 1999. Detection of infectious Cryptospori-
dium parvum oocysts in surface and ®lter backwash water samples by
immunomagnetic separation and integrated cell culture-PCR. Appl.
Environ. Microbiol. 65, 3427±32.
Doran, D.J., 1971. Increasing the yield of Eimeria tenella oocysts in cell
culture. J. Parasitol. 57, 891±900.
Dvorak, J., Crane, M.S.J., 1981. Vertebrate cell cycle modulates infection
by protozoan parasites. Science 214, 1034.
Dvorak, J., Howe, C., 1977. Toxoplasma gondii±vertebrate cell interaction.
I. The in¯uence of bicarbonate ion, CO
2
, pH and host cell culture age on
the invasion of vertebrate cells in vitro. J. Protozool. 24, 416.
Eggleston, M.T., Tilley, M., Upton, S.J., 1994. Enhanced development of
Cryptosporidium parvum in vitro by removal of oocyst toxins from
infected cell monolayers. Proc. Helm. Soc. Wash. 61, 118±21.
Kamamoto, F., Alejo-Blanco, R., Fleck, S.L., Sinden, R.E., 1991. Plasmo-
dium berghei: Ionic regulation and the induction of gametogenesis.
Exp. Parasitol. 72, 33±42.
Lawton, P., Naciri, M., Mancassola, R., Petavy, A., 1997. In vitro cultiva-
tion of Cryptosporidium parvum in the Non-Adherent Human Mono-
cytic THP-1 cell line. J. Eukar. Microbiol. 6, 44.
Meloni, B.P., Thompson, R.C.A., 1996. Simpli®ed method for obtaining
puri®ed oocysts from mice and growing Cryptosporidium parvum in
vitro. J. Parasitol. 82, 757±62.
Morgan, U.M., Constantine, C.C., Donoghue, P.O., Meloni, B.P., O'Brien,
P.A., Thompson, R.C.A., 1995. Molecular characterisation of Cryptos-
poridium isolates from humans and other animals using RAPD (random
ampli®ed polymorphic DNA) analysis. Am. J. Trop. Med. Hyg. 52,
559±64.
Morgan, U.M., Constantine, C.C., Forbes, D.A., Thompson, R.C.A., 1997.
Differentiation between human and animal isolates of Cryptosporidium
parvum using rDNA sequencing and direct PCR analysis. J. Parasitol.
83, 825±30.
Morgan, U.M., Pallant, L., Dwyer, B., Forbes, D.A., Thompson, R.C.A.,
1998. Comparison of PCR and microscopy for the detection of Cryp-
tosporidium in human specimens: clinical trial. J. Clin. Microbiol. 36,
995±8.
Morgan, U.M., Xiao, L., Fayer, R., Lal, A., Thompson, R.C.A., 1999.
Variation in Cryptosporidium: towards a taxonomic revision of the
genus. Int. J. Parasitol. 29, 1733±51.
Morgan, U.M., Weber, R., Xiao, L., Thompson, R.C.A., Ndiritu, W., Lal,
A., Moore, A., Deplazes, P., 2000. Molecular characterisation of Cryp-
tosporidium isolates obtained from human immunode®ciency virus-
infected individuals living in Switzerland, Kenya and the United States.
J. Clin. Microbiol. 38, 1180±3.
O'Donoghue, P.J., 1995. Cryptosporidium and Cryptosporidiosis in man
and animals. Int. J. Parasitol. 85, 525±30.
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±10551054
Peng, M.M., Xiao, L., Freeman, A.R., 1997. Genetic polymorphism among
Cryptosporidium parvum isolates: evidence of two distinct human
transmission cycles. Emerg. Infect. Dis. 3, 567±73.
Peterson, C., 1992. Cryptosporidiosis in patients with human immunode®-
ciency virus. Clin. Infect. Dis. 15, 903±9.
Pozio, E., Morales, M.A.G., Barbieri, F.M., LaRosa, G., 1992. Cryptospor-
idium: different behaviour in calves of isolates of human origin. Trans.
R. Soc. Trop. Med. Hyg. 86, 636±8.
Rochelle, P.A., De Leon, R., Johnson, A., Stewart, M.H., Wolfe, R.L.,
1999. Evaluation of immunomagnetic separation for the recovery of
infectious Cryptosporidium parvum oocysts from environmental
samples. Appl. Environ. Microbiol. 65, 841±5.
Sulaiman, I.M., Morgan, U.M., Thompson, R.C.A., Lal, A.A., Xiao, L.,
2000. Phylogenetic relationships of Cryptosporidium parasites based
on the 70-kilodalton heat shock protein (HSP70) gene. Appl. Environ.
Microbiol. 66, 2385±91.
Theodos, C.M., Grif®ths, J.K., D'onfro, J., Fair®eld, A., Tzipori, S., 1998.
Ef®cacy of nitazoxanide against Cryptosporidium parvum in cell
culture and in animal models. Antimicrob. Agents Chemother. 42,
1959±65.
Tzipori, S., 1998. Cryptosporidiosis: laboratory investigation and
chemotherapy. Adv. Parasitol. 40, 187±221.
Upton, S.J., 1997. In Vitro Cultivation. Cryptosporidiosis of Human and
Animals. CRC Press, Boca Raton, FL, pp. 181±207.
Upton, S.J., Tilley, M., Brillhart, D.B., 1995. Effects of select supplements
on the in vitro development of Cryptosporidium parvum in HCT-8 cells.
J. Clin. Microbiol. 33, 371±5.
Villacorta, I., Graaf, D.D., Charlier, G., Peeters, J., 1996. Complete devel-
opment of Cryptosporidium parvum in MDBK cells. FEMS Microbiol.
Lett. 142, 129±32.
Widmer, G., Tizipori, S., Fichtenbaum, C.J., Grif®ths, J.K., 1998. Genotyp-
ing and phenotyping characterisation of Cryptosporidium parvum
isolates from people with AIDS. J. Infect. Dis. 178, 834±40.
Widmer, G., Akiyoshi, D., Buckholt, M.A., Feng, X., Rich, S.M., Deary,
K.M., Bowman, C.A., Xu, P., Wang, X., Buch, G.A., Tzipori, S., 2000.
Animal propagation and genomic survey of genotype 1 isolate of Cryp-
tosporidium parvum. Mol. Biochem. Parasitol. 108, 187±97.
Xiao, L., Morgan, U.M., Limor, J., Escalante, A., Arrowood, M., Shulaw,
W., Thompson, R.C.A., Fayer, R., Lal, A.A., 1999. Genetic diversity
within Cryptosporidium parvum and related Cryptosporidium species.
Appl. Environ. Microbiol. 65, 3386±91.
Yang, S., Healey, M.C., Du, C., Zhang, J., 1996. Complete development of
Cryptosporidium parvum in bovine fallopian tube epithelial cells.
Infect. Immun. 64, 349±54.
N.S. Hijjawi et al. / International Journal for Parasitology 31 (2001) 1048±1055 1055
... Since the first report of Cryptosporidium cultivation in cell culture (Current and Haynes 1984), great strides have been made in the development and application of tools to both genetically manipulate Cryptosporidium and study its life cycle in vitro (Aldeyarbi and Karanis 2016;Borowski et al. 2010;Cardenas et al. 2020;Edwinson et al. 2016;English et al. 2022;Hashim et al. 2004;Hashim et al. 2006;Heo et al. 2018;Hijjawi et al. 2010;Hijjawi et al. 2001;Huang et al. 2004;Jumani et al. 2019;Mauzy et al. 2012;Miller et al. 2018;Morada et al. 2016;Pawlowic et al. 2019;Petry et al. 2009;Tandel et al. 2019;Tandel et al. 2023;Varughese et al. 2014;Vinayak et al. 2015;Warren et al. 2008;Wilke et al. 2019). The vast majority of these studies have focused on C. parvum due to its wide host range and readily available commercial stocks from animal sources (e.g. from Waterborne Inc, BioPoint Pty Ltd). ...
... The plates were incubated for a further 19 h (for 24-h infections) and 43 h (for 48-h infections), with infection medium refreshed at 24-h post-infection for plates designated for 48-h infections. These two time-points were selected specifically to assess the transition from asexual to sexual development which has been documented to occur in this time-frame (English et al. 2022;Hijjawi et al. 2001;Tandel et al. 2019). ...
... (Hashim et al. 2004). In contrast, other studies have reported minimal differences in the life cycle of C. parvum and C. hominis in HCT-8 cells, with the exception that C. hominis completed its life cycle more rapidly than C. parvum (72 h vs. 5 days) (Hijjawi et al. 2001). In the present study, there were no significant differences in the monolayer infection patterns between C. hominis and C. parvum, but clear differences in life cycle progression were observed. ...
Article
Full-text available
Cryptosporidium is a major cause of diarrhoeal disease and mortality in young children in resource-poor countries, for which no vaccines or adequate therapeutic options are available. Infection in humans is primarily caused by two species: C. hominis and C. parvum. Despite C. hominis being the dominant species infecting humans in most countries, very little is known about its growth characteristics and life cycle in vitro, given that the majority of our knowledge of the in vitro development of Cryptosporidium has been based on C. parvum. In the present study, the growth and development of two C. parvum isolates (subtypes Iowa-IIaA17G2R1 and IIaA18G3R1) and one C. hominis isolate (subtype IdA15G1) in HCT-8 cells were examined and compared at 24 h and 48 h using morphological data acquired with scanning electron microscopy. Our data indicated no significant differences in the proportion of meronts or merozoites between species or subtypes at either time-point. Sexual development was observed at the 48-h time-point across both species through observations of both microgamonts and macrogamonts, with a higher frequency of macrogamont observations in C. hominis (IdA15G1) cultures at 48-h post-infection compared to both C. parvum subtypes. This corresponded to differences in the proportion of trophozoites observed at the same time point. No differences in proportion of microgamonts were observed between the three subtypes, which were rarely observed across all cultures. In summary, our data indicate that asexual development of C. hominis is similar to that of C. parvum, while sexual development is accelerated in C. hominis. This study provides new insights into differences in the in vitro growth characteristics of C. hominis when compared to C. parvum, which will facilitate our understanding of the sexual development of both species.
... Human ileocaecal adenocarcinoma cells, originally obtained from American Type Culture Collection (ATCC; CCL-244™), Manassas, Virginia, were used for the in vitro maintenance of C. parvum infections and cultured in 25 cm 2 sterile polystyrene ventilated cell culture flasks (Cellstar ® ) in growth media consisting of RPMI 1640 (Sigma-Aldrich ® ) medium (10.3 g/L) supplemented with 10% foetal calf serum (FCS) (Sigma-Aldrich ® ) and other supplements as previously described by Hijjawi et al. [27]. The flasks were kept in a humidified incubator at 37 • C and 5% (v/v) CO 2 and grown for 24 h until monolayers reached 90% confluence. ...
... The oocysts were incubated for 30 min at 37 • C with vigorous shaking every 5 min. The excystation suspension was then centrifuged (at 2000× g for 8-10 min) and resuspended in the appropriate volume of maintenance media composed of RPMI 1640 (Sigma-Aldrich ® ) medium, 1% FCS and other supplements as described previously [27]. ...
Article
Full-text available
Cryptosporidium parvum is a significant cause of watery diarrhoea in humans and other animals worldwide. Although hundreds of novel drugs have been evaluated, no effective specific chemotherapeutic intervention for C. parvum has been reported. There has been much recent interest in evaluating plant-derived products in the fight against gastrointestinal parasites, including C. parvum. This study aimed to identify extracts from 13 different plant species that provide evidence for inhibiting the growth of C. parvum in vitro. Efficacy against C. parvum was detected and quantified using quantitative PCR and immunofluorescence assays. All plant extracts tested against C. parvum showed varying inhibition activities in vitro, and none of them produced a cytotoxic effect on HCT-8 cells at concentrations up to 500 µg/mL. Four plant species with the strongest evidence of activity against C. parvum were Curcuma longa, Piper nigrum, Embelia ribes, and Nigella sativa, all with dose-dependent efficacy. To the authors' knowledge, this is the first time that these plant extracts have proven to be experimentally efficacious against C. parvum. These results support further exploration of these plants and their compounds as possible treatments for Cryptosporidium infections.
... However, It does not support the production of new oocysts, suggesting a block to fertilization and zygote formation 6 . Alternative methods using HCT8s in a candle jar or COLO-680N esophageal cancer cells were reported to support production of new oocysts and, in the latter case, sustained oocyst production for up to 8 weeks 8,9 . Additionally, hollow nanofibers lined with HCT8s produced infectious oocysts for up to 20 weeks and maintained oocyst production for 2 years, but this has not been widely adopted and is not compatible with microscopy methods 10,11 . ...
Preprint
Full-text available
Typical cancer cell-based culture systems cannot support the full life cycle of Cryptosporidium parvum, despite its monoxenous life cycle which is completed in the small intestine of a single host. There is a block to fertilization and zygote formation in vitro. In this paper, we adapted a 2D organoid derived monolayer system and a 3D inverted enteroid system for use in C. parvum culture. 3D inverted enteroids were successfully infected by C. parvum without the need for microinjection and supported subculture of C. parvum. Using the 2D organoid derived monolayer (ODM) system, the infection can be maintained for at least 3 weeks with new oocyst production throughout. Fertilization was confirmed based on successful mating of two strains of C. parvum. We demonstrated that the apparent block to fertilization in typical cell culture is overcome using ODMs.
... The life cycle of C. mortiferum does not differ from that of previously described Cryptosporidium species. Consistent with studies that have examined the life cycle in vivo, we found no evidence for the occurrence of extracellular developmental stages described in cell-free cultures [57][58][59]. ...
Article
Full-text available
Background Cryptosporidium spp. are globally distributed parasites that infect epithelial cells in the microvillus border of the gastrointestinal tract of all classes of vertebrates. Cryptosporidium chipmunk genotype I is a common parasite in North American tree squirrels. It was introduced into Europe with eastern gray squirrels and poses an infection risk to native European squirrel species, for which infection is fatal. In this study, the biology and genetic variability of different isolates of chipmunk genotype I were investigated. Methods The genetic diversity of Cryptosporidium chipmunk genotype I was analyzed by PCR/sequencing of the SSU rRNA, actin, HSP70, COWP, TRAP-C1 and gp60 genes. The biology of chipmunk genotype I, including oocyst size, localization of the life cycle stages and pathology, was examined by light and electron microscopy and histology. Infectivity to Eurasian red squirrels and eastern gray squirrels was verified experimentally. Results Phylogenic analyses at studied genes revealed that chipmunk genotype I is genetically distinct from other Cryptosporidium spp. No detectable infection occurred in chickens and guinea pigs experimentally inoculated with chipmunk genotype I, while in laboratory mice, ferrets, gerbils, Eurasian red squirrels and eastern gray squirrels, oocyst shedding began between 4 and 11 days post infection. While infection in mice, gerbils, ferrets and eastern gray squirrels was asymptomatic or had mild clinical signs, Eurasian red squirrels developed severe cryptosporidiosis that resulted in host death. The rapid onset of clinical signs characterized by severe diarrhea, apathy, loss of appetite and subsequent death of the individual may explain the sporadic occurrence of this Cryptosporidium in field studies and its concurrent spread in the population of native European squirrels. Oocysts obtained from a naturally infected human, the original inoculum, were 5.64 × 5.37 μm and did not differ in size from oocysts obtained from experimentally infected hosts. Cryptosporidium chipmunk genotype I infection was localized exclusively in the cecum and anterior part of the colon. Conclusions Based on these differences in genetics, host specificity and pathogenicity, we propose the name Cryptosporidium mortiferum n. sp. for this parasite previously known as Cryptosporidium chipmunk genotype I. Graphical Abstract
... The main cell lines that have been used for the culture of the parasite are the human colorectal adenocarcinoma cell line (Caco-2) and the human ileocecal colorectal adenocarcinoma cell line . While some success has been found in the complete and long-term cultivation of the parasite in these cell lines (Hijjawi et al., 2001;Winkworth et al., 2008;Tandel et al., 2019), other studies have limited success in recreating these results and achieved partial progression of the life cycle terminating in the asexual phase . For this reason, more complex experimental designs, such as the use of hollow fiber technology to augment cell culture and other organoids and bioengineered intestinal models, have been proposed and implemented (Morada et al., 2016;Gunasekera et al., 2020). ...
Article
Full-text available
Cryptosporidiosis is a worldwide diarrheal disease caused by the protozoan Cryptosporidium. The primary symptom is diarrhea, but patients may exhibit different symptoms based on the species of the Cryptosporidium parasite they are infected with. Furthermore, some genotypes within species are more transmissible and apparently virulent than others. The mechanisms underpinning these differences are not understood, and an effective in vitro system for Cryptosporidium culture would help advance our understanding of these differences. Using COLO-680N cells, we employed flow cytometry and microscopy along with the C. parvum-specific antibody Sporo-Glo™ to characterize infected cells 48 h following an infection with C. parvum or C. hominis. The Cryptosporidium parvum-infected cells showed higher levels of signal using Sporo-Glo™ than C. hominis-infected cells, which was likely because Sporo-Glo™ was generated against C. parvum. We found a subset of cells from infected cultures that expressed a novel, dose-dependent auto-fluorescent signal that was detectable across a range of wavelengths. The population of cells that expressed this signal increased proportionately to the multiplicity of infection. The spectral cytometry results confirmed that the signature of this subset of host cells closely matched that of oocysts present in the infectious ecosystem, pointing to a parasitic origin. Present in both C. parvum and C. hominis cultures, we named this Sig M, and due to its distinct profile in cells from both infections, it could be a better marker for assessing Cryptosporidium infection in COLO-680N cells than Sporo-Glo™. We also noted Sig M’s impact on Sporo-Glo™ detection as Sporo-Glo™ uses fluoroscein–isothiocynate, which is detected where Sig M also fluoresces. Lastly, we used NanoString nCounter® analysis to investigate the transcriptomic landscape for the two Cryptosporidium species, assessing the gene expression of 144 host and parasite genes. Despite the host gene expression being at high levels, the levels of putative intracellular Cryptosporidium gene expression were low, with no significant difference from controls, which could be, in part, explained by the abundance of uninfected cells present as determined by both Sporo-Glo™ and Sig M analyses. This study shows for the first time that a natural auto-fluorescent signal, Sig M, linked to Cryptosporidium infection can be detected in infected host cells without any fluorescent labeling strategies and that the COLO-680N cell line and spectral cytometry could be useful tools to advance the understanding of Cryptosporidium infectivity.
... The lifecycle of Cryptosporidium can be replicated using cell culture in the laboratory, and the different parasite stages can be exposed to various compounds in vitro so that the comparative growth and proliferation may be determined (Hijjawi et al., 2001;Shahiduzzaman et al., 2009;Teichmann et al., 2012). Using this method, we here assessed the anti-Cryptosporidium potential of purified CTs obtained from a panel of five different plant species with CT purities >90% against zoonotic C. parvum oocysts. ...
Article
Full-text available
Infections with Cryptosporidium spp. Constitute a substantial public health burden and are responsible for widespread production losses in cattle herds. Reducing disease and shedding of Cryptosporidium oocysts is an important One Health goal. There are very few therapeutic options available to treat cryptosporidiosis. Interest in plant bioactive compounds to mitigate the spread of anthelmintic resistance in ruminants has led to investigation of these phytocompounds against other parasitic taxa. Condensed tannins (CTs) are plant secondary metabolites that have shown potential against nematodes in vitro and in vivo but their applicability to Cryptosporidium spp. Is comparatively under-explored. Cryptosporidium parvum infected human ileocecal colorectal adenocarcinoma (HCT)-8 cell cultures were treated with escalating doses of highly purified and well-characterized CTs from five plant species, big trefoil (Lotus pedunculatus), black currant (Ribes nigrum), sainfoin (Onobrychis viciifolia), white clover (Trifolium repens) and grapeseed (Vitis vinifera) for 44 h. Quantitative-PCR (qPCR) analysis revealed that none of the CTs examined demonstrated inhibitory potential against the parasite. Substantial inhibition of C. parvum by paromomycin was observed in positive controls in all assays (76.94–90.72% inhibition), proving the validity of the assay. Despite the lack of inhibition, these results represent an important step towards identifying alternative treatment options against this parasite.
... Likewise, in HCT-8 cells, a block of C. parvum development at the fertilization step was demonstrated [6]. In principle, developmental achievements in C. parvum in vitro propagation were reported for several cell lines, such as bovine fallopian tube epithelial cells BFTE, human Caco-2 cells, MBDK, HCT-8, and HFL [7][8][9]. Although those cell lines seem to support the full life cycle of Cryptosporidium, the yield of freshly produced infectious oocysts was minimal and insufficient for reinfection experiments. ...
Article
Full-text available
Cryptosporidium parvum is an important diarrhoea-associated protozoan, which is difficult to propagate in vitro. In 2017, a report described a continuous culture of C. parvum Moredun strain, in the oesophageal squamous cell carcinoma cell line COLO-680N, as an easy-to-use system for C. parvum propagation and continuous production of oocysts. Here, we report that—using the Köllitsch strain of C. parvum—even though COLO-680N cells, indeed, allowed parasite invasion and early asexual parasite replication, C. parvum proliferation decreased after the second day post infection. Considering recurring studies, reporting on successful production of newly generated Cryptosporidium oocysts in the past, and the subsequent replication failure by other research groups, the current data stand as a reminder of the importance of reproducibility of in vitro systems in cryptosporidiosis research. This is of special importance since it will only be possible to develop promising strategies to fight cryptosporidiosis and its ominous consequences for both human and animal health by a continuous and reliable methodological progress.
Article
Cryptosporidium spp. can cause diarrhea and even death in humans and animals. Host microRNAs (miRNAs) play an important role in the post-transcriptional regulation of the innate immune response to Cryptosporidium infection. To study host miRNA activity in the innate immune response to C. parvum infection, we examined the expression of miR-181d in HCT-8 cells infected with C. parvum and found that it was significantly downregulated, while TLR2, TLR4, NF-κB, and myD88 involved in the TLR/NF-κB signaling pathway were significantly upregulated at the early stages of C. parvum infection. We transfected cells with short-interfering RNAs (siRNA) as TLR2, TLR4, and NF-κB inhibitors. Analysis by quantitative real-time polymerase chain reaction (qPCR) and western blot confirmed that C. parvum downregulates miR-181d expression via the p50 subunit-dependent TLR2/TLR4-NF-κB signaling pathway in HCT-8 cells. This study provides a new theoretical foundation to elucidate the regulatory mechanism of host miRNAs against Cryptosporidium infection.
Article
Parasites belonging to the Apicomplexa phylum are among the most successful pathogens known in nature. They can infect a wide range of hosts, often remain undetected by the immune system, and cause acute and chronic illness. In this phylum, we can find parasites of human and veterinary health relevance, such as Toxoplasma, Plasmodium, Cryptosporidium, and Eimeria. There are still many unknowns about the biology of these pathogens due to the ethical and practical issues of performing research in their natural hosts. Animal models are often difficult or nonexistent, and as a result, there are apicomplexan life cycle stages that have not been studied. One recent alternative has been the use of three-dimensional (3D) systems such as organoids, 3D scaffolds with different matrices, microfluidic devices, organs-on-a-chip, and other tissue culture models. These 3D systems have facilitated and expanded the research of apicomplexans, allowing us to explore life stages that were previously out of reach and experimental procedures that were practically impossible to perform in animal models. Human- and animal-derived 3D systems can be obtained from different organs, allowing us to model host-pathogen interactions for diagnostic methods and vaccine development, drug testing, exploratory biology, and other applications. In this review, we summarize the most recent advances in the use of 3D systems applied to apicomplexans. We show the wide array of strategies that have been successfully used so far and apply them to explore other organisms that have been less studied.
Chapter
According to the World Health Organisation, cryptosporidiosis is a global diarrhoeal disease affecting millions of individuals; it is the second most common cause of infantile death in developing countries and is increasingly identified as an emerging cause of morbidity and mortality worldwide. The disease is also extremely severe in livestock, causing profuse diarrhoea and considerable economic losses in farmed young animals. Given the lack of effective treatment (absence of vaccines and effective drugs) and the limited understanding of the causative parasite, cryptosporidiosis represents a major challenge in the battle against global diarrhoeal diseases. Currently, there are 45 described Cryptosporidium species infecting a whole spectrum of animals. In this book chapter we will present an overview of the parasite, focusing on its taxonomic status, its morphology, its prevalence and transmission. We will review both cell biological and molecular techniques currently used to investigate the biology of this parasite and we will introduce the new state-of-the-art techniques that have been established by several laboratories in the field. With the development of these new technologies, we will be able to further understand the unique biology of Cryptosporidium and its role in health and disease of its host.
Article
Full-text available
PCR technology offers alternatives to conventional diagnosis of Cryptosporidium for both clinical and environmental samples. We compared microscopic examination by a conventional acid-fast staining procedure with a recently developed PCR test that can not only detect Cryptosporidium but is also able to differentiate between what appear to be host-adapted genotypes of the parasite. Examinations were performed on 511 stool specimens referred for screening on the basis of diarrhea. PCR detected a total of 36 positives out of the 511 samples, while routine microscopy detected 29 positives. Additional positives detected by PCR were eventually confirmed to be positive by microscopy. A total of five samples that were positive by routine microscopy at Western Diagnostic Pathology but negative by PCR and by microscopy in our laboratory were treated as false positives. Microscopy therefore exhibited 83.7% sensitivity and 98.9% specificity compared to PCR. PCR was more sensitive and easier to interpret but required more hands-on time to perform and was more expensive than microscopy. PCR, however, was very adaptable to batch analysis, reducing the costs considerably. Bulk buying of reagents and modifications to the procedure would decrease the cost of the PCR test even more. An important advantage of the PCR test, its ability to directly differentiate between different Cryptosporidium genotypes, will assist in determining the source of cryptosporidial outbreaks. Sensitivity, specificity, ability to genotype, ease of use, and adaptability to batch testing make PCR a useful tool for future diagnosis and studies on the molecular epidemiology of Cryptosporidium infections.
Article
Cryptosporidium is an important cause of enteric disease in humans and other animals. Limitations associated with conventional diagnostic methods for cryptosporidiosis based on morphological features, coupled with the difficulty of characterising parasites isolated in the laboratory, have restricted our ability to clearly identify species. The application of sensitive molecular approaches has obviated the necessity for laboratory amplification. Such studies have found considerable evidence of genetic heterogeneity among isolates of Cryptosporidium from different species of vertebrate, and there is now mounting evidence suggesting that a series of host-adapted genotypes/strains/species of the parasite exist. In this article, studies on the molecular characterisation of Cryptosporidium during the last 5 years are reviewed and put into perspective with the past and present taxonomy of the genus. The predictive value of achieving a sound taxonomy for the genus Cryptosporidium with respect to understanding its epidemiology and transmission and controlling outbreaks of the disease is also discussed.
Article
Much progress has been achieved in the last decade in terms of development of laboratory techniques, reagents and in vivo models. They have undoubtedly contributed to better and more accurate investigations. Despite a concerted effort by many investigators, however, breakthroughs have been minimal. The development of adequate in vitro and in vivo techniques for drug screening, and the intensified and systematic screening, has so far not resulted in the discovery of an effective therapy. The reason for the failure may well be due to the unique biological niche the parasite occupies (discussed at length in the first chapter in this volume). Its location beneath the cell membrane, but outside the cell cytoplasm, may prove a crucial element that needs to be considered when designing new therapeutic approaches. Laboratory investigations on two drugs currently used against chronic Cryptosporidium parvum in acquired immune deficiency syndrome (AIDS) are discussed. This chapter also provides information and the rationale for work in progress in our laboratory that relates to the development of novel approaches for control of the disease. This includes the identification of molecular targets of parasite origin for drug design, and studies on the structure-activity relationships of partially effective drugs with a view to synthesize more effective derivatives. Other investigations attempt to establish the role of secretory antibody, and the merit of repeated mucosal immunizations as a means of providing protection to individuals with AIDS who are at risk of developing chronic C. parvum infection.
Article
This study shows that the human monocytic cell line THP-1 supports the growth of C. parvum. Immunofluorescence controls showed that only scarce oocysts remained after infection and disappeared within the first 24 h of culture. A continuous asexual life cycle proceeded throughout the experiments, with at least 15-d cultures. This model provides a useful tool for studies on the biology of C. parvum in cells involved in its transport in immunocompromised host.
Article
A controlled-environment culture system was used to show that both physical and biologic parameters can influence the penetration of vertebrate cells by Toxoplasma gondii. The optimum bicarbonate ion concentration for the penetration of bovine embryo skeletal muscle (BESM) cells is 36.25 mM. Higher or lower bicarbonate ion concentrations are increasingly inhibitory to penetration. As CO2 increases in the range from 0.5-3.7 mM, penetration is progressively inhibited. No relationship was found between penetration and pH in the pH range of 6.949-7.765. The culture age of the BESM cells directly influenced the ability of the parasites to penetrate the cells. Older BESM cells were more refractory to penetration than younger cells.
Article
The behaviour in calves of 3 Cryptosporidium human isolates was analysed in comparison with a bovine isolate. Twenty-four neonatal calves were infected. An isolate from a patient infected with human immunodeficiency virus (HIV) and showing mild cryptosporidiosis caused severe diarrhoea with a high production of oocysts in neonatal calves, as did a bovine isolate (group 1). Two human isolates, obtained from HIV patients with severe cryptosporidiosis, caused mild diarrhoea with low oocyst production in neonatal calves (group 2). The difference between the 2 groups in numbers of oocysts shed in calves was statistically significant (P = 0.005), as was the duration of oocyst shedding (P = 0.0004). Oocysts of group 2 isolates were less resistant to storage in 2% potassium dichromate at 4 degrees C than were oocysts of group 1. The biological and epidemiological implications are discussed.
Article
We used the spontaneously differentiated human intestinal epithelial cell line Caco-2 to develop an in vitro model of Cryptosporidium sp. infection. The mean cell infection rate was 3% +/- 2%. Asexual stages of cryptosporidia were observed on day 2 postinoculation. Transmission electron microscopy showed the presence of macrogametes at day 5. This cell line appears to be suited to the study of the mechanisms by which biological agents inhibit both sexual and asexual development of cryptosporidia.
Article
The role of ionic regulation in the induction of gametogenesis of Plasmodium berghei at 20 degrees C was investigated. A potent inhibitor of Na+/H+ exchange, amiloride, strongly inhibited exflagellation and subsequent ookinete formation induced by RPMI 1640 with 10% fetal calf serum at pH 8.0, whereas Na+ or K+ channel inhibitors, H(+)-ATPase inhibitors, and a protonophore had no significant effect. Amiloride-treated 'activated' microgametocytes synthesized DNA to levels consistent with the expected 8C, but failed to develop further. These results may suggest that an increase in intracellular pH induced by Na+/H+ exchange plays an important role in the induction of gametogenesis by cultivating at pH 8.0 and 20 degrees C. Cultivation at pH 8.0 and 37 degrees C did not induce the development, and microgametocytes remained as nonactivated forms, having the DNA content of 1.5C. By culturing at pH 7.3 and 20 degrees C, however, most of microgametocytes finished synthesis of DNA up to the 8C level, but ceased development at various stages. Additionally, exflagellation occurred in a simple medium composed of buffered saline with 10 mM glucose. Glucose was indispensable for exflagellation, presumably acting as an energy source. Exflagellation induced by this solution was also inhibited by amiloride. It is therefore suggested that the induction of microgametogenesis may be composed of two distinct mechanisms, one is a temperature-dependent DNA synthesis and the other is a pH-dependent control of developmental events leading to microgamete assembly and exflagellation.
Article
Suckling mice were used as a model host to compare the endogenous development of three different isolates of Cryptosporidium: one from a naturally infected calf, one from an immunocompetent human with a short-term diarrheal illness, and one from a patient with acquired immune deficiency syndrome (AIDS) and persistent, life-threatening, gastrointestinal cryptosporidiosis. After oral inoculation of mice with oocysts, no differences were noted among developmental stages of the three isolates in their sites of infection, times of appearance, and duration, morphology, and fine structure. Sporozoites excysted within the lumen of the duodenum and ileum, penetrated into the microvillous region of villous enterocytes, and developed into type I meronts with six or eight merozoites. Type I merozoites penetrated enterocytes and underwent cyclic development as type I meronts or they became type II meronts with four merozoites. Type II merozoites did not exhibit cyclic development but developed directly into sexual forms. Microgamonts produced approximately 16 small, bullet-shaped microgametes, which were observed attaching to and penetrating macrogametes. Approximately 80% of the oocysts observed in enterocytes had a thick, two-layered wall. After sporulating within the parasitophorous vacuole, these thick-walled oocysts passed through the gut unaltered and were the resistant forms that transmitted the infection to a new host. Approximately 20% of the oocysts in enterocytes consisted of four sporozoites and a residuum surrounded only by a single oocyst membrane that ruptured soon after the parasite was released from the host cell. The presence of thin-walled, autoinfective oocysts and recycling of type I meronts may explain why a small oral inoculum can produce an overwhelming infection in a suitable host and why immune deficient persons can have persistent, life-threatening cryptosporidiosis in the absence of repeated oral exposure to thick-walled oocysts.