Open access peer-reviewed chapter

State of the Art and Future Directions of Cryptosporidium spp.

Written By

Helena Lúcia Carneiro Santos, Karina Mastropasqua Rebello and Teresa Cristina Bergamo Bomfim

Submitted: 02 April 2019 Reviewed: 03 July 2019 Published: 10 September 2019

DOI: 10.5772/intechopen.88387

From the Edited Volume

Parasitology and Microbiology Research

Edited by Gilberto Antonio Bastidas Pacheco and Asghar Ali Kamboh

Chapter metrics overview

1,041 Chapter Downloads

View Full Metrics

Abstract

Cryptosporidium species are protozoan parasites that infect epithelium surfaces in gastrointestinal and respiratory tracts of humans and a range of animals worldwide. Cryptosporidiosis has been associated with considerable morbidity and, under certain circumstances, mortality. Humans can acquire it by consuming food and drink containing oocysts, which have been recognised as a major cause for diarrhoeal disease. The ubiquitousness of the infective oocyst, its resilience to environmental pressures, and the low dose of oocyst exposure needed for infection amplify to outbreaks of Cryptosporidium traced to drinking and recreational water. Unlike in developing countries where lack of sustained access to safe water creates tremendous burdens of Cryptosporidium diarrhoea, this scenario is aggravated due to limited diagnosis and therapeutics. However, over the past few decades, growing information on Cryptosporidium genomes have allowed novel insight into the host-parasite relationship. Future field research on potential tools will focus on biology-derived parasite products applicable to drugs and diagnosis. This chapter reviews available data on biology, transmission, life cycle, diagnosis, genome, and a few but important progresses in the field of cryptosporidiosis.

Keywords

  • cryptosporidiosis
  • diagnosis
  • transmission
  • infectious disease
  • genome

1. Introduction

Cryptosporidium species are protozoan parasites that infect the epithelial cells of the gastrointestinal and respiratory tracts of humans and a wide range of animals, with a global distribution [1, 2, 3]. Cryptosporidium represents a major public health concern for waterborne disease and daycare outbreaks of diarrhoeal disease worldwide [2, 4, 5, 6, 7, 8]. Human cryptosporidiosis is usually a self-limiting infection in immunocompetent individuals. However, cases of severe diarrhoea and dissemination to extra-intestinal sites can occur in children, the elderly, and individuals with impairment of T-cell functions, mainly those with HIV infection [9, 10, 11, 12]. In children, although diarrhoea is a key feature of malabsorption, it may not be apparent at presentation; when the infection becomes chronic, the only symptom may be limited growth. Consequently, chronic infections can culminate in poor growth [5, 13, 14, 15, 16]. The epidemiology of infections is complex and involves transmission by a faecal-oral route, either by ingestion of contaminated water or food or by human-to-human or animal-to-human transmission [17, 18]. The oocyst, the environmental stage of Cryptosporidium, is incredibly hardy, easily spread through water, and resistant to inactivation by chlorine; and without the use of filtration, it is challenging to remove it from drinking water [19, 20, 21]. Cryptosporidium prevalence is higher in areas lacking a sanitation infrastructure, mainly drinking water and sewage, which led the World Health Organization (WHO) to include it in the water sanitation and health programme [22]. The scarcity of sustained access to safe water creates tremendous burdens of Cryptosporidium diarrhoea in developing countries [23]. Treatment and diagnosis options are still not totally effective [2, 24, 25, 26]. No fully effective drug therapy or vaccine is available for Cryptosporidium, and the diagnosis of cryptosporidiosis has been based on the demonstration of oocysts in faeces, which present low sensibility [25]. However, the ability to culture relevant Cryptosporidium isolates in vitro, the development of novel gene-editing tools (knockout genes, CRISPR/Cas9, and RNAi) [26, 27, 28, 29, 30], and ‘omic’ research (genomics, transcriptomics, and proteomics) represent essential paths towards significant advancements in the control of cryptosporidiosis [30, 31, 32, 33, 34, 35, 36, 37, 38]. In the future, those approaches will show a holistic view of the biology of Cryptosporidium. In this chapter, we present recent advances and remaining challenges regarding human cryptosporidiosis under a public health perspective.

Advertisement

2. Clinical perspective, diagnosis, and treatment

Despite Cryptosporidium species infecting the epithelial cells of the gastrointestinal and respiratory epithelium tracts, human cryptosporidiosis is a usually self-limiting infection in immunocompetent individuals with a low fatality rate [39, 40, 41]. In general, onset of the symptoms occurs 5–7 days following exposure and resolves in 2–3 weeks [42]. Clinical manifestations vary from subclinical infection to watery diarrhoea, sometimes profuse. Other common symptoms include abdominal cramps, fever, flatulence, nausea, vomiting, and low-grade fever [43, 44, 45]. Clinical presentation of cryptosporidiosis in individuals with impairment of T-cell functions, mainly those with HIV infection, varies according to the level of immunosuppression, from asymptomatic disease, to transient disease, to relapsing chronic diarrhoea or even cholera-like diarrhoea that is debilitating and potentially life-threatening [46]. Spreading of infection beyond the extra-intestinal site (in the biliary or respiratory tract) has been documented in children and immunocompromised people, resulting in a potentially life-threatening disease [47, 48]. Sclerosing cholangitis and other biliary involvements are common in AIDS patients with cryptosporidiosis. Both innate and adaptive immunity of the host have major impacts on the severity of cryptosporidiosis and its prognosis.

Cryptosporidium has been diagnosed using a variety of approaches, such as microscopy, immunofluorescent antibody (IFA), enzyme-linked immunosorbent assay (ELISA), and DNA-based detection methods [18]. However, identification of the parasite’s morphologic features through examination of stool smears is widely employed in diagnostic laboratories, particularly in resource-limited health systems. The oocysts are shed intermittently [49]; therefore, three faecal samples collected on alternate days are recommended. To maximise the recovery of oocysts, Sheather’s sucrose flotation, saturated salt flotation, and Allen and Ridley’s formol-ether method are the stool concentration techniques most frequently used prior to the use of the microscopy staining technique [50, 51]. Stain differential is required due to the small size of the specimen (ranging from 4 to 6 μm), similar in shape to yeasts and faecal debris [52]. Safranin-methylene blue, Kinyoun Ziehl-Neelsen, and dimethyl sulfoxide-carbol fuchsin are the most commonly used stain methods [11, 53, 54, 55]. However, in the absence of staining solution, phase contrast microscopy has proven to be highly specific for the detection of Cryptosporidium oocysts in human stool samples [56]. In general, conventional microscopy lacks sensitivity, is time-consuming, and requires a skilled and well-experienced microscopist [57, 58, 59].

Direct fluorescent antibody tests (DFAs), enzyme-immunoassays (EIAs), and rapid immunochromatographic assays (dipsticks) are commercially available [60, 61, 62, 63]. The EIA kits have been evaluated with human stool specimens only, presumably from patients infected with C. hominis or C. parvum. The direct fluorescent antibody tests have been widely used for the detection of Cryptosporidium in faecal smears, water, and food [60, 62, 63, 64, 65, 66]. However, the antigenic variability of oocyst wall epitopes contributes to reducing specificity, and the sensibility of all immunological-based methods is low. High specificity (99–100%) has been generally reported for EIA kits. Sensitivities, however, have been reported to range from 70 to 100% [62, 63, 64, 65]. Dipsticks and EIAs are available for individual and for all-in-one tests for Giardia, Cryptosporidium, and Entamoeba histolytica [66, 67, 68, 69]. The tests are fast and easy to perform. However, EIA kits and rapid format assays present a potential problem with false positives, so results need to be interpreted and evaluated with caution [70]. To overcome these barriers, one of the most notable advances in public health in recent decades has been the development of tools based on molecular biology for the diagnosis of infectious diseases. These polymerase chain reaction (PCR) techniques have enabled specific sensitive detection of oocysts (a single oocyst) in clinical and environmental samples [71, 72, 73, 74, 75, 76, 77]. Examples of such techniques include conventional PCR, quantitative PCR real time, and high-resolution melt. A wide variety of PCR methods targeting different genes have been developed for the detection of Cryptosporidium at the species/genotype/subtype levels. However, no targeted tests have been patterned for the detection of Cryptosporidium in clinical laboratories. Recently, the simultaneous qualitative detection and identification of multiple viral, parasitic (including C. parvum and C. hominis), and bacterial nucleic acids in human stool specimens were approved by the Food and Drug Administration (FDA) [78]. In general, PCR tools solely amplify the DNA of C. parvum, C. hominis, C. meleagridis, and species/genotypes closely related to C. parvum [18]. For genotyping, nested PCR-RFLP was the most commonly used method in the past. Nowadays, DNA sequencing of 18S has been required to reliably detect all Cryptosporidium spp. The HSP70 and COWP targets fail to detect the DNA of C. felis, C. canis, and C. muris [79]. Subtyping tools are indispensable from the epidemiological point of view and are helpful in knowing the possible transmission routes of Cryptosporidium species and zoonotic potential of the parasite. Several subtyping tools have been developed to evaluate the diversity within C. parvum or C. hominis, including analysis of the microsatellite, GP-60 gene, HSP70 gene, 47-kDa protein, small double-stranded (ds) RNA virus, serine repeat antigen, and T-rich gene fragment [73, 80, 81, 82, 83, 84, 85]. The 18S ribosomal RNA (rRNA) gene and the hypervariable 60-kDa glycoprotein (gp60) gene have been widely used as targets to identify species and track transmission [18, 86, 87]. The 60-kDa glycoprotein (gp60, also known as gp40/gp15) gene presents a wide genetic heterogeneity in the number of trinucleotide repeats (TCA, TCG, or TCT). This gene encodes a precursor protein that is cleaved to produce mature cell surface glycoproteins (gp45/gp40 and gp15) implicated in the attachment to, and invasion of, enterocytes [18, 87]. Identification of subtypes using GP60 subtype families has revealed the subtype families (Ia-Ik) in C. hominis [87, 88, 89, 90, 91] and two zoonotic subtypes (IIa, IId), subtypes (IIb, IIc, IIe, IIf, IIi, IIj-IIt) in C. parvum [4, 87, 92, 93, 94], and subtype families (IIIa to IIIg) in C. meleagridis have been acknowledged [87, 95, 96]. Subtyping tools targeting the gp60 gene have been developed recently for several other human-pathogenic Cryptosporidium species [87]. Species and subtype identification are not necessary for clinical care and therapeutic options but are important for epidemiological surveillance and for drug investigations and clinical trials. Novel diagnostic tools and biomarkers for cryptosporidiosis, which could also be used for therapeutic or vaccine trials, are necessary for accurate identification.

Current treatment options for cryptosporidiosis are limited. So far, there is no vaccine against Cryptosporidium [97], and nitazoxanide (NTZ) is the only drug approved by the FDA for treatment of cryptosporidiosis in children and immunocompetent adults [98]. However, it is not effective without an appropriate immune status and, consequently, is ineffectual for the treatment of immune-compromised patients, particularly those with AIDS [25, 99]. NTZ is a nitrothiazole benzamide compound with a broad spectrum of activity against a wide range of parasites, bacteria, and viruses. In protozoa, NTZ inhibits the enzyme pyruvate ferredoxin oxidoreductase, which is essential to anaerobic energy metabolism [100]. Due to the prevalence of cryptosporidiosis, the development of novel therapeutic targets and vaccines against Cryptosporidium spp. is a public health priority. The ongoing need to develop new anti-cryptosporidial drugs has spurred the process of finding new uses for existing drugs. Repurposing drug provides an attractive alternative to drug development [101]. Two compounds, 3-hydroxy-3-methyl-glutaryl-coenzyme A (HMG-CoA) reductase inhibitor, pitavastatin and auranofin (approved for the treatment of rheumatoid arthritis), have been shown to be effective against Cryptosporidium in vitro [102]. Auranofin has been shown to be 10 times more potent than metronidazole against Entamoeba histolytica, the protozoan agent of human amoebiasis [103]. HMG-CoA and auranofin have particular promise in fast-tracking for further in vivo testing in animals and humans.

Advertisement

3. Life cycle and classification

The parasite has a complex monoxenous life cycle with both asexual (merogony) and sexual (gametogony) stages. Ingestion of an infective oocyst (containing four sporozoites) by a susceptible host initiates the excystation process in the gastrointestinal tract. The sporulated oocyst ruptures, releasing sporozoites that invade the enterocytes, inducing the cell membrane to enclose the parasite in the parasitophorous vacuole, which then differentiates into a trophozoite. Trophozoites undergo merogony and form either a further type I meront or a type II meront, which contains four merozoites that are destined for gametogony. Merozoites can differentiate into sexually distinct stages called macro- and microgametocytes in a process called gametogony. New oocysts are formed in the epithelial cells from the fusion of a macro- and a microgametocyte to form a diploid zygote. The new fused cell evolves and sporulates in situ in a process called sporogony, becoming oocysts containing four sporozoites. Type II meronts attach to the epithelial cell and differentiate into either macrogamonts or microgamonts. The microgametes from the microgamont are released, and each can fertilised a macrogamont to form a diploid zygote. This cell undergoes a process like meiosis (sporogony) to produce an oocyst, either thin- or thick-walled, containing four sporozoites (sporulated oocysts). The thin-walled oocysts are involved in autoinfection, and thick-walled oocysts are released within the faeces to infect new hosts [104, 105, 106, 107] (Figure 1).

Figure 1.

A schematic diagram of Cryptosporidium life cycle. After ingestion of contaminated water and/or food, the oocyst wall opens (excystation) triggered by temperature, stomach acid, and bile salts. Then, sporulated oocyst ruptures releasing (a) sporozoites that (b) invade the host cell (c) inducing the cell membrane to enclose the parasite in the parasitophorous vacuole, (d) which then differentiates into a trophozoite that undergoes an asexual reproduction, (e) forming a type I meront that contains 6–8 merozoites. These merozoites can reinfect the epithelial cell, where they undergo merogony and form type I meront or (f) type II meront. (g) Merozoites can differentiate into sexually distinct stages called (g) micro- and (h) macrogametocytes. (i) Zygote is formed after the fertilisation of macrogametocyte by the microgametocyte, (j) and this cell undergoes sporogony and produces a thin-walled oocyst. (k) These thin-walled oocysts are released within faeces to infect new hosts, as well as (l) involved in autoinfection process (adapted from Ref. [104]).

Until relatively recently, Cryptosporidium was classified as a coccidian parasite. However, the taxonomic placement of Cryptosporidium was altered after revisions to higher-order classifications due to recent particularities observed in Cryptosporidium. The parasite can develop in a cell-free culture, while extracellular stages have been observed in both cell-free and cell cultures, in biofilms, and in vivo [108, 109, 110, 111]. It presents the ability to grow and amplify without host cell attachment and encapsulation, as well as the insensitivity of all anticoccidial agents [26]. Moreover, the parasite lacks a micropyle, sporocyst, and polar granular [111, 112, 113]. Although initially considered to be a coccidian, Cryptosporidium spp. share features of both the coccidia and gregarines, confirmed by morphological and molecular data. Major similarities between Cryptosporidium and gregarine parasites are as follows: (1) the ability to complete its life cycle in the absence of host cells, (2) extracellular gamont-like stages, (3) the process in which two mature trophozoites pair up before the formation of gametocyst (szygy), and (4) changing cell architecture to adapt to diverse environments (biofilms, coelom, intestines, soil, and water) [107, 108, 111, 114]. The most recent classification considers Cryptosporidium as a separate group within the Apicomplexa. Analyses of comparative genomics and of phylogenetic inference and the ability of Cryptosporidium to complete its life cycle extracellularly confirm its close relationship with gregarines and corroborate the transference of Cryptosporidium to the Gregarinomorphea class as a new subclass of Cryptogregaria [111, 115]. Early taxonomy at species level was based originally on morphology and host specificity. Nowadays, the description of species takes molecular analyses, mainly DNA sequencing and PCR-related methods, into account for the detection and differentiation of Cryptosporidium spp.

Advertisement

4. Maintenance of Cryptosporidium in nature and transmission

Once excreted into the environment, oocysts can be dispersed from the faecal matrix into the terrestrial environment (Figure 2). When present on the soil surface, oocysts may be exposed to high temperatures and desiccation, causing their inactivation. Oocysts are sensitive to desiccation and UV-C irradiation [116]. Reports show that desiccation is lethal to oocysts with only 3 and 5% remaining viable after being air-dried at room temperature for 2 and 4 h, respectively [117, 118]. However, when within the soil column, the oocysts were maintained, protected, and viable [119, 120]. Studies have indicated that oocysts at 4°C recovered from soil column may remain infectious for long periods [119, 121]. These findings suggest that the soil column is a sanctuary for Cryptosporidium, protecting it until rainfall events scatter them [120]. Oocysts were able to remain viable and infectious after being frozen at −10°C for up to 168 h, at −15°C for up to 24 h, and at −20°C for up to 8 h [122]. Moreover, Cryptosporidium oocysts can be carried in the environment due to interactions with biofilms (surface-attached microbial communities). They readily attach to biofilms and persist and subsequently separate from it. High concentrations of oocysts in water biofilms that were maintained over several months maintained viable sporozoites [123]. Cryptosporidium oocysts in fresh water and marine water can survive at a range of temperatures. Fayer et al. reported that oocysts maintained at 20°C remain infectious for 12 weeks at salinities of 0 and 10 ppt, for 4 weeks at 20 ppt, and for 2 weeks at 30 ppt [124]. Although salinity can have a pronounced effect on oocyst infectivity, they can survive long enough in marine waters to justify their presence in marine animals.

Figure 2.

Cryptosporidium sp. oocysts in safranin-methylene blue staining method.

Cryptosporidium spp. have a huge impact on both human and veterinary health worldwide, aggravated by the limited diagnosis and current therapeutics. Cryptosporidium spp. have a worldwide distribution and the ability to infect a wide range of hosts, including humans, and a broad variety of vertebrate [1, 3]. Humans can acquire cryptosporidiosis through several transmission routes, such as direct contact with infected persons or animals and consumption of contaminated water (drinking or recreational) or food (Figure 3).

Figure 3.

Transmission cycles of Cryptosporidium infections.

The WHO has categorised Cryptosporidium as a reference pathogen for the assessment of drinking water quality [125]. Susceptibility to cryptosporidiosis depends on several factors, including environmental conditions, host immune status, age, geographic location, and contact with infected humans/animals [126]. Animals play an important role in the maintenance, amplification, and transmission of Cryptosporidium [127]. In fact, a large range of animals are reservoirs for some species, genotypes, and subtypes, which may infect humans [128, 129, 130]. The lack of adequate instruments to continuously monitor animal mobility makes it difficult to study the dynamics of transmission [131, 132]. Also, oocysts are ubiquitous in the environment and easily spread via drinking water, recreational water, and food [3, 133, 134]. The ubiquitousness of the infective oocyst, its resilience to environmental pressures [135], and the low-dose oocyst exposure (ingestion of fewer than 10 oocysts can lead to infection) [136, 137] amplify to outbreaks of Cryptosporidium traced to drinking and recreational water. In 1993, the largest Cryptosporidium waterborne outbreak was recorded in the United States in Milwaukee, where more than 400,000 people were infected by the drinking water supply [138]. The epidemiology of infection is complex and involves transmission by the faecal-oral route, either by indirect transmission through ingestion of contaminated water or food or by direct human-to-human or animal-to-human transmission [3]. The genus Cryptosporidium has about 30 species formally described, as well as various genotypes and subtypes. Some species are relatively promiscuous in terms of host specificity, some of which also infected humans. Currently, a wide range of Cryptosporidium species and various genotypes have been recognised as responsible for human cryptosporidiosis (Table 1).

Cryptosporidium spp. Major host References
C. andersoni Cattle [139, 140, 141, 142, 143, 144]
C. baileyi Chickens and turkeys [145]
C. bovis Cattle [146, 147]
C. canis Dogs [148, 149, 150]
C. cuniculus Rabbits [151, 152, 153, 154, 155]
C. fayeri Marsupials [141, 156]
C. felis Cats [93, 157, 158]
C. hominis * Humans [18]
C. meleagridis * Turkeys, chickens,
humans
[18, 93, 153, 159, 160]
C. muris Rodents [161, 162, 163]
C. parvum * ruminants, especially calves [18]
C. scrofarum Pigs [164]
C. suis Pigs [139, 160, 165, 166, 167]
C. tyzzeri Rodents, snake [168]
C. ubiquitum Sheep and cervids [152, 154, 157, 158, 169]
C. viatorum Humans [93, 170]
C. erinacei Hedgehogs and horses [171, 172]
C. wrairi Guinea pigs [173, 174]
C. xiaoi Sheep and goat [174]
Cryptosporidium
Chipmunk genotype
Rodents [93]
Cryptosporidium
Horse genotype
Horses [152, 153]
Cryptosporidium
Mink genotype
Minks [175]
Cryptosporidium
Monkey genotype
Monkey [152]
Cryptosporidium
Skunk genotype
Skunk [152, 153]

Table 1.

Currently recognised species of Cryptosporidium spp. associated with human infections.

The most prevalent species.


Human infections predominantly are caused by C. hominis, which are considered restricted to humans (anthroponotic transmission), and by C. parvum, some of which isolate genotypes and infect ruminants (zoonotic transmission) [18]. However, in recent years, C. meleagridis, C. cuniculus, and C. ubiquitum have also emerged as species relevant to public health, while the other species tend to be associated only with sporadic and rare cases of human infection. Approximately 155 species of mammals have been reported as non-human hosts of C. parvum, indicating that the parasite is adapting and developing in many hosts [176].

The human-to-human spread is particularly well described within families (often secondary cases after a primary outbreak infection) in childcare nurseries, nursing homes, and hospitals [42, 177, 178]. In addition, contact with production animals, mainly cattle, that are the main hosts of C. parvum can potentially infect humans [40, 178, 179]. To date, studies in developing countries have shown a predominance of C. hominis in HIV-positive children and adults. These findings are also valid in the United States, Canada, Australia, and Japan. In Europe and New Zealand, several studies have shown a similar prevalence of C. parvum and C. hominis in immunocompetent and immunocompromised individuals. Thus, in most developing countries, the anthroponotic transmission of Cryptosporidiumplays an important role in human cryptosporidiosis [18, 180], while in Europe, New Zealand, and rural areas of the United States, there are both anthroponotic and zoonotic transmissions. In Middle Eastern countries, children are mainly infected with C. parvum, but the significance of this occurrence is not clear [181]. An exception is Cryptosporidium infections in HIV-positive patients in Ababa, Ethiopia, where C. parvum is highly endemic and where contact with calves is an important risk factor for cryptosporidiosis [174].

In developing countries, most C. parvum infections in HIV-positive children and adults are caused by subtype IIc, with IIa largely absent, indicating that anthroponotic transmission of C. parvum is common in these areas. Conversely, families of subtype IIa are commonly diagnosed in humans in industrialised regions, where their occurrence is often associated with contact with calves. Another family of C. parvum subtypes commonly found in sheep and goats, IId, is dominant in humans in Middle Eastern countries and is occasionally found in humans in some European countries, such as Sweden, where it is commonly diagnosed in dairy calves. A systematic review of the anthroponotic transmission of Cryptosporidium concluded that subtype IIc predominates in low-income countries with poor sanitation and in HIV-positive individuals, unlike in higher-income countries, where it is rarely evident. Lacking effective treatment or vaccine, intervention to improve basic sanitation in these regions is the best option. This prophylactic action certainly may reduce the anthropogenic and zoonotic transmission of cryptosporidiosis, reducing the damage to human health. It is important to emphasise the importance of personal hygiene practices to minimise cryptosporidiosis, in addition to other pathogens transmitted by water and food.

Advertisement

5. Genome of Cryptosporidium: new insight and future challenges

Recent years have seen impressive progress of next-generation sequencing technologies in genome assembly and annotation methodologies, mainly by advancements in the fields of molecular biology and technical engineering and by reducing cost. Cryptosporidium has been the subject of genome sequencing projects, which have provided valuable insights into the species, biology, and host-parasite relationships. The genomic data of multiple Cryptosporidium species are available and accessible in a Cryptosporidium-dedicated database, CryptoDB (http://cryptodb.org/cryptodb/) [182], and in the GenBank database (www.ncbi.nlm.nih.gov). Comparative analyses have shown that Cryptosporidium genomes are highly compact, containing 8.50–9.50 megabase pairs (Mbp), a total gene count ranging from 3769 to 7610, and coding sequence composition (75–77.6%). Moreover, in general, they share a comparable GC percentage (Table 2).

Organism/name Strain Bio-sample Bio-project Size (Mb) GC % Gene Protein
C. hominis SAMEA 3496639 PRJEB 10000 9.10 30.1 3818 3817
C. hominis TU502 PRJNA 13200 8.74 30.9 3949 3885
C. hominis TU502_2012 SAMN 02382005 PRJNA 222836 9.10 30.1 3796 3745
C. hominis UKH1 SAMN 02382004 PRJNA 222837 9.15 30.1 3769 3718
C. hominis 30,976 SAMN 02862040 PRJNA 252787 9.06 30.1 3995 3959
C. parvum Iowa type II SAMN 02952908 PRJNA 144 9.10 30.2 7774 7610
C. andersoni 30,847 SAMN 04417240 PRJNA 354069 9.09 28.5 3897 3876
C. meleagridis UKMEL1 SAMN 02666797 PRJNA 222838 8.97 31.0 3806 3753
C. meleagridis UKMEL4 SAMN 08383028 PRJNA 315503 8.79 30.9
C. meleagridis UKMEL3 SAMN 08383027 PRJNA 315502 8.70 31.0
C. ubiquitum 39,726 SAMN 02768023 PRJNA 534291 8.97 30.8 3766 3766
Cryptosporidium sp. Chipmunk LX-2015 SAMN 03281121 PRJNA 272389 9.51 31.9
Cryptosporidium sp. 37,763 SAMN 10623052 PRJNA 511361 9.05 32.0
C. baileyi * TAMU 09Q1 SAMN 02382006 PRJNA 222835 8.50 24.2
C. cuniculus UKCU2 SAMN 08383019 PRJNA 3154496 9.18 25.8
C. muris RN66 SAMN 02953683 PRJA 19553 9.25 28.5 3934
C. viatorum UKUIA1 SAMN 10107889 PRJA 492837 9.26 31.1

Table 2.

Genomic features of Cryptosporidium spp.

Draft genome.


Overall, gene content and genomic organisation among intestinal occurrences of the species are well conserved, with Cryptosporidium gene clusters encoding putative secreted proteins. Comparison of the Cryptosporidium genomes has identified a core set of proteins commonly studied, as well as major differences in particular gene families, which could be involved in biological differences between species and genotypes [114, 183, 184, 185]. Gene encoding proteins that are associated with invasion processes, e.g. protein kinases and thrombospondin-related adhesive proteins (TRAPs), insulinase-like peptidases, MEDLE secretory proteins, and mucin glycoproteins, are observed in genome Cryptosporidium spp. [32]. However, some of them differ in copy number variations of genes. Comparative genomic analysis revealed that one of the primary features differentiating Cryptosporidium species is the sequence diversity present in major secreted protein families, MEDLE, and insulinase-like proteases [184]. This is consistent with transcriptomic studies of C. parvum, which have demonstrated MEDLE proteins in different subcellular locations that may perform their functions in distinct stages of the invasion and development process [33]. Moreover, a reduction in the number of genes encoding secreted MEDLE and insulinase-like proteins was observed in C. ubiquitum and C. andersoni, whereas the mucin-type glycoproteins are highly divergent between the gastric C. andersoni and intestinal Cryptosporidium species [184]. Unlike most other apicomplexans, Cryptosporidium spp. have no apicoplast or mitochondrial genomes but have remnant ones, the so-called mitosomes. However, Cryptosporidium species disagree from each other mostly in mitosome metabolic pathways. C. parvum, C. hominis, and C. andersoni present more aerobic metabolism and a conventional electron transport chain [114], whereas C. ubiquitum has further reductions in ubiquinone and polyisoprenoid biosynthesis and has lost both the conventional and alternative electron transport systems, unlike C. muris genome encoding core enzymes for the Krebs cycle and a functional ATP synthase. Thus, the mitosome of C. muris functions essentially as a peculiar mitochondrion [186]. However, the loss of biosynthetic pathways is a common feature observed in Cryptosporidium spp. genomes, e.g. the cytochrome-based respiratory chain and main de novo synthetic pathways for amino acids, nucleotides, fatty acids, and the Krebs cycle [32, 183]. Conversely, families of transporters to acquire nutrients from the host were expanded, including transporters for amino acids, sugars, and ATP-binding cassettes (ABCs) that drive the transport of various metabolites, lipids/sterols, and drugs [32]. Although these genomic sequences provide valuable data, the genome analyses have revealed contradictory data and inconsistencies between the annotated gene models and transcriptome evidence [31, 36]. Notoriously, those findings are related to sequencing platforms, which have been applied to having different strengths and weaknesses and the use of different strategies and stringencies in gene prediction.

Notwithstanding its novelty, the major challenges for the generation of whole genomes of Cryptosporidium are the quality and the yielding of limited DNA. Indeed, this is a critical step, as it is hard to recover enough quantity of DNA (2.5 × 10−5 highly purified oocysts correspondent approximately 10 μg) from faeces from natural infections. A theoretical estimate of the DNA content of one oocyst is of 40 fg [187]; therefore, it is tricky and arduous to recover enough quantity of DNA (2.5 × 10−5 highly purified oocysts correspondent approximately 10 μg) from unculturable samples with the quality necessary for high-throughput sequencing. Non-cultured samples may introduce a level of uncertainty and possess limited metadata. The lower the quality of the initial genome sequence, the higher the likelihood of yielding a missing or misassembled genome. A recent study evaluated an alternative method of preparing faecal samples using the combination of salt flotation, immunomagnetic separation (IMS), and surface sterilisation of oocysts prior to DNA extraction. The method has shown promise when used for the genome sequencing of samples of C. parvum and C. hominis [36]. This challenging issue can be resolved using a novel approach of Cryptosporidium cell-free culture and new long-read sequencing techniques, which will likely be beneficial for improving data. Increases in the quality of the target DNA boost the depth of coverage of the genome in higher levels, so base calls can be made with a higher degree of confidence. Also, the ability to culture relevant Cryptosporidium isolates in vitro, the development of novel gene-editing tools (knockout genes, CRISPR/Cas9, and RNAi), and ‘omic’ research (genomics, transcriptomics, and proteomics) represent essential paths towards significant advancements in the control of cryptosporidiosis.

Advertisement

6. Conclusions and future perspectives

Cryptosporidium is a major cause of diarrhoeal disease in humans worldwide, yet an effective therapy to eradicate the parasite is not available. Also, the diagnosis options remain limited in developing countries, which harm the surveillance and understanding of the epidemiology in resource-poor settings. In developed countries, large waterborne outbreaks in drinking and recreational water continue to occur, emphasising the need for better regulation and for improvements of drinking water treatment processes and control guidelines. However, in recent years, significant improvements have been achieved in understanding the key concepts of the organism, mainly by increasing the use of molecular methods and genome sequences. Recent advancements in knowledge of Cryptosporidium provide the basis for the development of effective and practical strategies for the future prevention and control of cryptosporidiosis. The data from Cryptosporidium genome sequences have already improved our understanding of the metabolism and cellular processes. In fact, mining the genome and proteome data of Cryptosporidium will allow the development of new classes of compounds and molecular targets. However, it is worth underscoring the need for community-wide efforts to generate and integrate high-quality functional datasets that span the full spectrum of biology and life cycles in order to improve the predictive nature of models generated from large-scale system-based resources. Transcriptomes and proteomics from different growth stages are starting to be generated and promise to provide further insight into the biology of Cryptosporidium. Also, future studies will require careful validation and follow-up of each finding using in vitro and animal model studies.

Advertisement

Acknowledgments

I would like to thank Victor Ricardo Azevedo for the assistance with Figure 3.

Advertisement

Conflict of interest

The authors declare that there is no conflict of interest.

References

  1. 1. Feng Y, Ryan UM, Xiao L. Genetic diversity and population structure of Cryptosporidium. Trends in Parasitology. 2018;34(11):997-1011. DOI: 10.1016/j.pt.2018.07.009. PubMed PMID: 30108020
  2. 2. Checkley W, White AC Jr, Jaganath D, Arrowood MJ, Chalmers RM, Chen XM, et al. A review of the global burden, novel diagnostics, therapeutics, and vaccine targets for Cryptosporidium. The Lancet Infectious Diseases. 2015;15(1):85-94. DOI: 10.1016/S1473-3099(14)70772-8. PubMed PMID: 25278220
  3. 3. Ryan U, Fayer R, Xiao L. Cryptosporidium species in humans and animals: Current understanding and research needs. Parasitology. 2014;141(13):1667-1685. DOI: 10.1017/S0031182014001085. PubMed PMID: 25111501
  4. 4. Ajjampur SS, Liakath FB, Kannan A, Rajendran P, Sarkar R, Moses PD, et al. Multisite study of cryptosporidiosis in children with diarrhea in India. Journal of Clinical Microbiology. 2010;48(6):2075-2081. DOI: 10.1128/JCM.02509-09. PubMed PMID: 20392919
  5. 5. Kotloff KL, Blackwelder WC, Nasrin D, Nataro JP, Farag TH, van Eijk A, et al. The global enteric multicenter study (GEMS) of diarrheal disease in infants and young children in developing countries: Epidemiologic and clinical methods of the case/control study. Clinical Infectious Diseases. 2012;55(Suppl 4):S232-S245. DOI: 10.1093/cid/cis753. PubMed PMID: 23169936
  6. 6. Shirley DA, Moonah SN, Kotloff KL. Burden of disease from cryptosporidiosis. Current Opinion in Infectious Diseases. 2012;25(5):555-563. DOI: 10.1097/QCO.0b013e328357e569. PubMed PMID: 22907279
  7. 7. Korpe PS, Valencia C, Haque R, Mahfuz M, McGrath M, Houpt E, et al. Epidemiology and risk factors for cryptosporidiosis in children from 8 low-income sites: Results from the MAL-ED study. Clinical Infectious Diseases. 2018;67(11):1660-1669. DOI: 10.1093/cid/ciy355. PubMed PMID: 29701852
  8. 8. Wheeler C, Vugia DJ, Thomas G, Beach MJ, Carnes S, Maier T, et al. Outbreak of cryptosporidiosis at a California waterpark: Employee and patron roles and the long road towards prevention. Epidemiology and Infection. 2007;135(2):302-310. DOI: 10.1017/S0950268806006777. PubMed PMID: 17291365
  9. 9. Baldursson S, Karanis P. Waterborne transmission of protozoan parasites: Review of worldwide outbreaks-An update 2004-2010. Water Research. 2011;45(20):6603-6614. DOI: 10.1016/j.watres.2011.10.013. PubMed PMID: 22048017
  10. 10. Navin TR, Juranek DD. Cryptosporidiosis: Clinical, epidemiologic, and parasitologic review. Reviews of Infectious Diseases. 1984;6(3):313-327. PubMed PMID: 6377439
  11. 11. Ma P, Soave R. Three-step stool examination for cryptosporidiosis in 10 homosexual men with protracted watery diarrhea. The Journal of Infectious Diseases. 1983;147(5):824-828. DOI: 10.1093/infdis/147.5.824. PubMed PMID: 6842020
  12. 12. Guerrant DI, Moore SR, Lima AA, Patrick PD, Schorling JB, Guerrant RL. Association of early childhood diarrhea and cryptosporidiosis with impaired physical fitness and cognitive function four-seven years later in a poor urban community in Northeast Brazil. The American Journal of Tropical Medicine and Hygiene. 1999;61(5):707-713. DOI: 10.4269/ajtmh.1999.61.707. PubMed PMID: 10586898
  13. 13. Operario DJ, Platts-Mills JA, Nadan S, Page N, Seheri M, Mphahlele J, et al. Etiology of severe acute watery diarrhea in children in the global rotavirus surveillance network using quantitative polymerase chain reaction. The Journal of Infectious Diseases. 2017;216(2):220-227. DOI: 10.1093/infdis/jix294. PubMed PMID: 28838152
  14. 14. Krause I, Amir J, Cleper R, Dagan A, Behor J, Samra Z, et al. Cryptosporidiosis in children following solid organ transplantation. The Pediatric Infectious Disease Journal. 2012;31(11):1135-1138. DOI: 10.1097/INF.0b013e31826780f7. PubMed PMID: 22810017
  15. 15. Wang RJ, Li JQ , Chen YC, Zhang LX, Xiao LH. Widespread occurrence of Cryptosporidium infections in patients with HIV/AIDS: Epidemiology, clinical feature, diagnosis, and therapy. Acta Tropica. 2018;187:257-263. DOI: 10.1016/j.actatropica.2018.08.018. PubMed PMID: 30118699
  16. 16. Liu L, Johnson HL, Cousens S, Perin J, Scott S, Lawn JE, et al. Global, regional, and national causes of child mortality: An updated systematic analysis for 2010 with time trends since 2000. Lancet. 2012;379(9832):2151-2161. DOI: 10.1016/S0140-6736(12)60560-1. PubMed PMID: 22579125
  17. 17. Caccio SM, Chalmers RM. Human cryptosporidiosis in Europe. Clinical Microbiology and Infection. 2016;22(6):471-480. DOI: 10.1016/j.cmi.2016.04.021. PubMed PMID: 27172805
  18. 18. Xiao L. Molecular epidemiology of cryptosporidiosis: An update. Experimental Parasitology. 2010;124(1):80-89. DOI: 10.1016/j.exppara.2009.03.018. PubMed PMID: 19358845
  19. 19. Carpenter C, Fayer R, Trout J, Beach MJ. Chlorine disinfection of recreational water for Cryptosporidium parvum. Emerging Infectious Diseases. 1999;5(4):579-584. DOI: 10.3201/eid0504.990425. PubMed PMID: 10458969
  20. 20. Water Safety Plans: Managing Drinking Water Quality from Catchment to Consumer [Internet]. Geneva: World Health Organization; 2005
  21. 21. Pollock KG, Young D, Robertson C, Ahmed S, Ramsay CN. Reduction in cryptosporidiosis associated with introduction of enhanced filtration of drinking water at loch Katrine. Scotland. Epidemiol Infect. 2014;142(1):56-62. DOI: 10.1017/S0950268813000678. PubMed PMID: 23591075
  22. 22. WHO. Risk Assessment of Cryptosporidium in Drinking-Water Guidelines for Drinking-Water Quality. Geneva: WHO; 2011. pp. 303-304
  23. 23. Manjunatha UH, Chao AT, Leong FJ, Diagana TT. Cryptosporidiosis drug discovery: Opportunities and challenges. ACS Infectious Diseases. 2016;2(8):530-537. DOI: 10.1021/acsinfecdis.6b00094. PubMed PMID: 27626293
  24. 24. Ryan U, Paparini A, Oskam C. New technologies for detection of enteric parasites. Trends in Parasitology. 2017;33(7):532-546. DOI: 10.1016/j.pt.2017.03.005. PubMed PMID: 28385423
  25. 25. Amadi B, Mwiya M, Sianongo S, Payne L, Watuka A, Katubulushi M, et al. High dose prolonged treatment with nitazoxanide is not effective for cryptosporidiosis in HIV positive Zambian children: A randomised controlled trial. BMC Infectious Diseases. 2009;9:195. DOI: 10.1186/1471-2334-9-195. PubMed PMID: 19954529
  26. 26. Cabada MM, White AC Jr. Treatment of cryptosporidiosis: Do we know what we think we know? Current Opinion in Infectious Diseases. 2010;23(5):494-499. DOI: 10.1097/QCO.0b013e32833de052. PubMed PMID: 20689422
  27. 27. Vinayak S, Pawlowic MC, Sateriale A, Brooks CF, Studstill CJ, Bar-Peled Y, et al. Genetic modification of the diarrhoeal pathogen Cryptosporidium parvum. Nature. 2015;523(7561):477-480. DOI: 10.1038/nature14651. PubMed PMID: 26176919
  28. 28. Beverley SM. Parasitology: CRISPR for Cryptosporidium. Nature. 2015;523(7561):413-414. DOI: 10.1038/nature14636. PubMed PMID: 26176915
  29. 29. Witola WH, Zhang X, Kim CY. Targeted gene knockdown validates the essential role of lactate dehydrogenase in Cryptosporidium parvum. International Journal for Parasitology. 2017;47(13):867-874. DOI: 10.1016/j.ijpara.2017.05.002. PubMed PMID: 28606696
  30. 30. Castellanos-Gonzalez A, Perry N, Nava S, White AC. Preassembled single-stranded RNA-Argonaute complexes: A novel method to silence genes in Cryptosporidium. The Journal of Infectious Diseases. 2016;213(8):1307-1314. DOI: 10.1093/infdis/jiv588. PubMed PMID: 26656125
  31. 31. Ifeonu OO, Chibucos MC, Orvis J, Su Q , Elwin K, Guo F, et al. Annotated draft genome sequences of three species of Cryptosporidium: Cryptosporidium meleagridis isolate UKMEL1, C. baileyi isolate TAMU-09Q1 and C. hominis isolates TU502_2012 and UKH1. Pathogens and Disease. 2016;74(7):1-5. DOI: 10.1093/femspd/ftw080. PubMed PMID: 27519257
  32. 32. Xu Z, Guo Y, Roellig DM, Feng Y, Xiao L. Comparative analysis reveals conservation in genome organization among intestinal Cryptosporidium species and sequence divergence in potential secreted pathogenesis determinants among major human-infecting species. BMC Genomics. 2019;20(1):406. DOI: 10.1186/s12864-019-5788-9. PubMed PMID: 31117941
  33. 33. Su J, Jin C, Wu H, Fei J, Li N, Guo Y, et al. Differential expression of three Cryptosporidium species-specific MEDLE proteins. Frontiers in Microbiology. 2019;10:1177. DOI: 10.3389/fmicb.2019.01177. PubMed PMID: 31191495
  34. 34. Widmer G. Diverse single-amino-acid repeat profiles in the genus Cryptosporidium. Parasitology. 2018;145(9):1151-1160. DOI: 10.1017/S0031182018000112. PubMed PMID: 29429420
  35. 35. Nader JL, Mathers TC, Ward BJ, Pachebat JA, Swain MT, Robinson G, et al. Evolutionary genomics of anthroponosis in Cryptosporidium. Nature Microbiology. 2019;4(5):826-836. DOI: 10.1038/s41564-019-0377-x. PubMed PMID: 30833731
  36. 36. Hadfield SJ, Pachebat JA, Swain MT, Robinson G, Cameron SJ, Alexander J, et al. Generation of whole genome sequences of new Cryptosporidium hominis and Cryptosporidium parvum isolates directly from stool samples. BMC Genomics. 2015;16:650. DOI: 10.1186/s12864-015-1805-9. PubMed PMID: 26318339
  37. 37. Zhang H, Guo F, Zhou H, Zhu G. Transcriptome analysis reveals unique metabolic features in the Cryptosporidium parvum oocysts associated with environmental survival and stresses. BMC Genomics. 2012;13:647. DOI: 10.1186/1471-2164-13-647. PubMed PMID:23171372
  38. 38. Xu P, Widmer G, Wang Y, Ozaki LS, Alves JM, Serrano MG, et al. The genome of Cryptosporidium hominis. Nature. 2004;431(7012):1107-1112. DOI: 10.1038/nature02977. PubMed PMID: 15510150
  39. 39. Clark DP. New insights into human cryptosporidiosis. Clinical Microbiology Reviews. 1999;12(4):554-563. PubMed PMID: 10515902
  40. 40. Hunter PR, Hughes S, Woodhouse S, Syed Q , Verlander NQ , Chalmers RM, et al. Sporadic cryptosporidiosis case-control study with genotyping. Emerging Infectious Diseases. 2004;10(7):1241-1249. DOI: 10.3201/eid1007.030582. PubMed PMID: 15324544
  41. 41. Chalmers RM, Davies AP. Minireview: Clinical cryptosporidiosis. Experimental Parasitology. 2010;124(1):138-146. DOI: 10.1016/j.exppara.2009.02.003. PubMed PMID: 19545516
  42. 42. Hunter PR, Hadfield SJ, Wilkinson D, Lake IR, Harrison FC, Chalmers RM. Subtypes of Cryptosporidium parvum in humans and disease risk. Emerging Infectious Diseases. 2007;13(1):82-88. DOI: 10.3201/eid1301.060481. PubMed PMID: 17370519
  43. 43. Fayer R. Ungar BL, Cryptosporidium spp. and cryptosporidiosis. Microbiology Reviews. 1986;50(4):458-483. PubMed PMID: 3540573
  44. 44. Casemore DP. Epidemiological aspects of human cryptosporidiosis. Epidemiology and Infection. 1990;104(1):1-28. DOI: 10.1017/s0950268800054480. PubMed PMID: 2407541
  45. 45. Chappell CL, Okhuysen PC, Langer-Curry R, Widmer G, Akiyoshi DE, Tanriverdi S, et al. Cryptosporidium hominis: Experimental challenge of healthy adults. American Journal of Tropical Medicine and Hygiene. 2006;75(5):851-857. PubMed PMID: 17123976
  46. 46. Ryan U, Zahedi A, Paparini A. Cryptosporidium in humans and animals-a one health approach to prophylaxis. Parasite Immunology. 2006;38(9):535-547. DOI: 10.1111/pim.12350. PubMed PMID: 27454991
  47. 47. Mercado R, Buck GA, Manque PA, Ozaki LS. Cryptosporidium hominis infection of the human respiratory tract. Emerging Infectious Diseases. 2007;13(3):462-464. DOI: 10.3201/eid1303.060394. PubMed PMID: 17552101
  48. 48. Xiao L, Feng Y. Zoonotic cryptosporidiosis. FEMS Immunology and Medical Microbiology. 2008;52(3):309-323. DOI: 10.1111/j.1574-695X.2008.00377.x. PubMed PMID: 18205803
  49. 49. Vanathy K, Parija SC, Mandal J, Hamide A, Krishnamurthy S. Cryptosporidiosis: A mini review. Tropical Parasitology. 2017;7(2):72-80. DOI: 10.4103/tp.TP_25_17. PubMed PMID: 29114483
  50. 50. McNabb SJ, Hensel DM, Welch DF, Heijbel H, McKee GL, Istre GR. Comparison of sedimentation and flotation techniques for identification of Cryptosporidium sp. oocysts in a large outbreak of human diarrhea. Journal of Clinical Microbiology. 1985;22(4):587-589. PubMed PMID: 2416771
  51. 51. Alles AJ, Waldron MA, Sierra LS, Mattia AR. Prospective comparison of direct immunofluorescence and conventional staining methods for detection of Giardia and Cryptosporidium spp. in human fecal specimens. Journal of Clinical Microbiology. 1995;33(6):1632-1634. PubMed PMID: 7544365
  52. 52. O’Donoghue PJ. Cryptosporidium and cryptosporidiosis in man and animals. International Journal for Parasitology. 1995;25(2):139-195. PubMed PMID: 7622324
  53. 53. Baxby D, Blundell N, Hart CA. The development and performance of a simple, sensitive method for the detection of Cryptosporidium oocysts in faeces. The Journal of Hygiene. 1984;93(2):317-323. DOI: 10.1017/s0022172400064858. PubMed PMID: 6209333
  54. 54. Pohjola S, Jokipii L, Jokipii AM. Dimethylsulphoxide-Ziehl-Neelsen staining technique for detection of cryptosporidial oocysts. The Veterinary Record. 1985;116(16):442-443. PubMed PMID: 2408372
  55. 55. Henriksen SA, Pohlenz JF. Staining of cryptosporidia by a modified Ziehl-Neelsen technique. Acta Veterinaria Scandinavica. 1981;22(3-4):594-596. PubMed PMID: 6178277
  56. 56. Ignatius R, Klemm T, Zander S, Gahutu JB, Kimmig P, Mockenhaupt FP, et al. Highly specific detection of Cryptosporidium spp. oocysts in human stool samples by undemanding and inexpensive phase contrast microscopy. Parasitology Research. 2016;115(3):1229-1234. DOI: 10.1007/s00436-015-4859-3. PubMed PMID: 26646397
  57. 57. Smith HV, McDiarmid A, Smith AL, Hinson AR, Gilmour RA. An analysis of staining methods for the detection of Cryptosporidium spp. oocysts in water-related samples. Parasitology. 1989;99(Pt 3):323-327. PubMed PMID: 2481834
  58. 58. Moodley D, Jackson TF, Gathiram V, van den Ende J. Cryptosporidium infections in children in Durban. Seasonal variation, age distribution and disease status. South African Medical Journal. 1991;79(6):295-297. PubMed PMID: 2017736
  59. 59. Fall A, Thompson RC, Hobbs RP, Morgan-Ryan U. Morphology is not a reliable tool for delineating species within Cryptosporidium. The Journal of Parasitology. 2003;89(2):399-402. DOI: 10.1645/0022-3395(2003)089[0399:MINART]2.0.CO;2. PubMed PMID: 12760666
  60. 60. Garcia LS, Bruckner DA, Brewer TC, Shimizu RY. Techniques for the recovery and identification of Cryptosporidium oocysts from stool specimens. Journal of Clinical Microbiology. 1983;18(1):185-190. PubMed PMID: 6193138
  61. 61. Geurden T, Thomas P, Casaert S, Vercruysse J, Claerebout E. Prevalence and molecular characterisation of Cryptosporidium and Giardia in lambs and goat kids in Belgium. Veterinary Parasitology. 2008;155(1-2):142-145. DOI: 10.1016/j.vetpar.2008.05.002. PubMed PMID: 18565678
  62. 62. Robinson TJ, Cebelinski EA, Taylor C, Smith KE. Evaluation of the positive predictive value of rapid assays used by clinical laboratories in Minnesota for the diagnosis of cryptosporidiosis. Clinical Infectious Diseases. 2010;50(8):e53-e55. DOI: 10.1086/651423. PubMed PMID: 20218890
  63. 63. Agnamey P, Sarfati C, Pinel C, Rabodoniriina M, Kapel N, Dutoit E, et al. Evaluation of four commercial rapid immunochromatographic assays for detection of Cryptosporidium antigens in stool samples: A blind multicenter trial. Journal of Clinical Microbiology. 2011;49(4):1605-1607. DOI: 10.1128/JCM.02074-10. PubMed PMID: 21289154
  64. 64. Garcia LS, Shimizu RY. Evaluation of nine immunoassay kits (enzyme immunoassay and direct fluorescence) for detection of Giardia lamblia and Cryptosporidium parvum in human fecal specimens. Journal of Clinical Microbiology. 1997;35(6):1526-1529. PubMed PMID: 9163474
  65. 65. Bialek R, Binder N, Dietz K, Joachim A, Knobloch J, Zelck UE. Comparison of fluorescence, antigen and PCR assays to detect Cryptosporidium parvum in fecal specimens. Diagnostic Microbiology and Infectious Disease. 2002;43(4):283-288. PubMed PMID: 12151188
  66. 66. Srijan A, Wongstitwilairoong B, Pitarangsi C, Serichantalergs O, Fukuda CD, Bodhidatta L, et al. Re-evaluation of commercially available enzyme-linked immunosorbent assay for the detection of Giardia lamblia and Cryptosporidium spp. from stool specimens. The Southeast Asian Journal of Tropical Medicine and Public Health. 2005;36(Suppl 4):26-29. PubMed PMID: 16438175
  67. 67. Chalmers RL, Wagner H, Mitchell GL, Lam DY, Kinoshita BT, Jansen ME, et al. Age and other risk factors for corneal infiltrative and inflammatory events in young soft contact lens wearers from the Contact Lens Assessment in Youth (CLAY) study. Investigative Ophthalmology & Visual Science. 2011;52(9):6690-6696. DOI: 10.1167/iovs.10-7018. PubMed PMID: 21527379
  68. 68. Llorente MT, Clavel A, Varea M, Olivera S, Castillo FJ, Sahagun J, et al. Evaluation of an immunochromatographic dip-strip test for the detection of Cryptosporidium oocysts in stool specimens. European Journal of Clinical Microbiology & Infectious Diseases. 2002;21(8):624-625. DOI: 10.1007/s10096-002-0778-1. PubMed PMID: 12226697
  69. 69. Weitzel T, Dittrich S, Mohl I, Adusu E, Jelinek T. Evaluation of seven commercial antigen detection tests for Giardia and Cryptosporidium in stool samples. Clinical Microbiology and Infection. 2006;12(7):656-659. DOI: 10.1111/j.1469-0691.2006.01457.x. PubMed PMID: 16774562
  70. 70. Chalmers RM, Campbell BM, Crouch N, Charlett A, Davies AP. Comparison of diagnostic sensitivity and specificity of seven Cryptosporidium assays used in the UK. Journal of Medical Microbiology. 2011;60(Pt 11):1598-1604. DOI: 10.1099/jmm.0.034181-0. PubMed PMID: 21757501
  71. 71. Soliman RH, Othman AA. Evaluation of DNA melting curve analysis real-time PCR for detection and differentiation of Cryptosporidium species. Parasitologists United Journal (PUJ). 2009;2(1):47-54
  72. 72. Mary C, Chapey E, Dutoit E, Guyot K, Hasseine L, Jeddi F, et al. Multicentric evaluation of a new real-time PCR assay for quantification of Cryptosporidium spp. and identification of Cryptosporidium parvum and Cryptosporidium hominis. Journal of Clinical Microbiology. 2013;51(8):2556-2563. DOI: 10.1128/JCM.03458-12. PubMed PMID: 23720792
  73. 73. Spano F, Putignani L, McLauchlin J, Casemore DP, Crisanti A. PCR-RFLP analysis of the Cryptosporidium oocyst wall protein (COWP) gene discriminates between C. wrairi and C. parvum, and between C. parvum isolates of human and animal origin. FEMS Microbiology Letters. 1997;150(2):209-217. DOI: 10.1016/s0378-1097(97)00115-8. PubMed PMID: 9170264
  74. 74. Abe N, Matsubayashi M, Kimata I, Iseki M. Subgenotype analysis of Cryptosporidium parvum isolates from humans and animals in Japan using the 60-kDa glycoprotein gene sequences. Parasitology Research. 2006;99(3):303-305. DOI: 10.1007/s00436-006-0140-0. PubMed PMID: 16565816
  75. 75. Jothikumar N, da Silva AJ, Moura I, Qvarnstrom Y, Hill VR. Detection and differentiation of Cryptosporidium hominis and Cryptosporidium parvum by dual TaqMan assays. Journal of Medical Microbiology. 2008;57(Pt 9):1099-1105. DOI: 10.1099/jmm.0.2008/001461-0. PubMed PMID: 18719179
  76. 76. Sturbaum GD, Reed C, Hoover PJ, Jost BH, Marshall MM, Sterling CR. Species-specific, nested PCR-restriction fragment length polymorphism detection of single Cryptosporidium parvum oocysts. Applied and Environmental Microbiology. 2001;67(6):2665-2668. DOI: 10.1128/AEM.67.6.2665-2668.2001. PubMed PMID: 11375178
  77. 77. Hadfield SJ, Robinson G, Elwin K, Chalmers RM. Detection and differentiation of Cryptosporidium spp. in human clinical samples by use of real-time PCR. Journal of Clinical Microbiology. 2011;49(3):918-924. DOI: 10.1128/JCM.01733-10. PubMed PMID: 21177904
  78. 78. Navidad JF, Griswold DJ, Gradus MS, Bhattacharyya S. Evaluation of Luminex xTAG gastrointestinal pathogen analyte-specific reagents for high-throughput, simultaneous detection of bacteria, viruses, and parasites of clinical and public health importance. Journal of Clinical Microbiology. 2013;51(9):3018-3024. DOI: 10.1128/JCM.00896-13. PubMed PMID: 23850948
  79. 79. Jiang J, Xiao L. An evaluation of molecular diagnostic tools for the detection and differentiation of human-pathogenic Cryptosporidium spp. The Journal of Eukaryotic Microbiology. 2003;50:542-547. PubMed PMID: 14736156
  80. 80. Robinson G, Chalmers RM. Assessment of polymorphic genetic markers for multi-locus typing of Cryptosporidium parvum and Cryptosporidium hominis. Experimental Parasitology. 2012;132(2):200-215. DOI: 10.1016/j.exppara.2012.06.016. PubMed PMID: 22781277
  81. 81. Sulaiman IM, Lal AA, Xiao L. Molecular phylogeny and evolutionary relationships of Cryptosporidium parasites at the actin locus. The Journal of Parasitology. 2002;88(2):388-394. DOI: 10.1645/0022-3395(2002)088[0388:MPAERO]2.0.CO;2. PubMed PMID: 12054017
  82. 82. Spano F, Putignani L, Crisanti A, Sallicandro P, Morgan UM, Le Blancq SM, et al. Multilocus genotypic analysis of Cryptosporidium parvum isolates from different hosts and geographical origins. Journal of Clinical Microbiology. 1998;36(11):3255-3259. PubMed PMID: 9774575
  83. 83. Pedraza-Diaz S, Amar C, McLauchlin J. The identification and characterisation of an unusual genotype of Cryptosporidium from human faeces as Cryptosporidium meleagridis. FEMS Microbiology Letters. 2000;189(2):189-194. DOI: 10.1111/j.1574-6968.2000.tb09228.x. PubMed PMID: 10930736
  84. 84. Feng Y, Yang W, Ryan U, Zhang L, Kvac M, Koudela B, et al. Development of a multilocus sequence tool for typing Cryptosporidium muris and Cryptosporidium andersoni. Journal of Clinical Microbiology. 2011;49(1):34-41. DOI: 10.1128/JCM.01329-10. PubMed PMID: 20980577
  85. 85. Yadav P, Mirdha BR, Makharia GK, Chaudhry R. Multilocus sequence typing of Cryptosporidium hominis from northern India. The Indian Journal of Medical Research. 2017;145(1):102-111. DOI: 10.4103/ijmr.IJMR_1064_14. PubMed PMID: 28574022
  86. 86. Plutzer J, Karanis P. Genetic polymorphism in Cryptosporidium species: An update. Veterinary Parasitology. 2009;165(3-4):187-199. DOI: 10.1016/j.vetpar.2009.07.003. PubMed PMID: 19660869
  87. 87. Xiao L, Feng Y. Molecular epidemiologic tools for waterborne pathogens Cryptosporidium spp. and Giardia duodenalis. Food and Waterborne Parasitology. 2017;8-9:14-32
  88. 88. Molloy SF, Smith HV, Kirwan P, Nichols RA, Asaolu SO, Connelly L, et al. Identification of a high diversity of Cryptosporidium species genotypes and subtypes in a pediatric population in Nigeria. The American Journal of Tropical Medicine and Hygiene. 2010;82(4):608-613. DOI: 10.4269/ajtmh.2010.09-0624. PubMed PMID: 20348508
  89. 89. Feng Y, Lal AA, Li N, Xiao L. Subtypes of Cryptosporidium spp. in mice and other small mammals. Experimental Parasitology. 2011;127(1):238-242. DOI: 10.1016/j.exppara.2010.08.002. PubMed PMID: 20692256
  90. 90. Li W, Kiulia NM, Mwenda JM, Nyachieo A, Taylor MB, Zhang X, et al. Cyclospora papionis, Cryptosporidium hominis, and human-pathogenic Enterocytozoon bieneusi in captive baboons in Kenya. Journal of Clinical Microbiology. 2011;49(12):4326-4329. DOI: 10.1128/JCM.05051-11. PubMed PMID: 21956988
  91. 91. Laatamna AE, Wagnerova P, Sak B, Kvetonova D, Xiao L, Rost M, et al. Microsporidia and Cryptosporidium in horses and donkeys in Algeria: Detection of a novel Cryptosporidium hominis subtype family (Ik) in a horse. Veterinary Parasitology. 2015;208(3-4):135-142. DOI: 10.1016/j.vetpar.2015.01.007. PubMed PMID: 25638716
  92. 92. Hira KG, Mackay MR, Hempstead AD, Ahmed S, Karim MM, O’Connor RM, et al. Genetic diversity of Cryptosporidium spp. from Bangladeshi children. Journal of Clinical Microbiology. 2011;49(6):2307-2310. DOI: 10.1128/JCM.00164-11. PubMed PMID: 21471344
  93. 93. Insulander M, Silverlas C, Lebbad M, Karlsson L, Mattsson JG, Svenungsson B. Molecular epidemiology and clinical manifestations of human cryptosporidiosis in Sweden. Epidemiology and Infection. 2013;141(5):1009-1020. DOI: 10.1017/S0950268812001665. PubMed PMID: 22877562
  94. 94. Liu X, Zhou X, Zhong Z, Zuo Z, Shi J, Wang Y, et al. Occurrence of novel and rare subtype families of Cryptosporidium in bamboo rats (Rhizomys sinensis) in China. Veterinary Parasitology. 2015;207(1-2):144-148. DOI: 10.1016/j.vetpar.2014.11.009. PubMed PMID: 25499825
  95. 95. Vermeulen ET, Ashworth DL, Eldridge MD, Power ML. Diversity of Cryptosporidium in brush-tailed rock-wallabies (Petrogale penicillata) managed within a species recovery programme. The International Journal for Parasitology: Parasites and Wildlife. 2015;4(2):190-196. DOI: 10.1016/j.ijppaw.2015.02.005. PubMed PMID: 25834789
  96. 96. Stensvold CR, Beser J, Axen C, Lebbad M. High applicability of a novel method for gp60-based subtyping of Cryptosporidium meleagridis. Journal of Clinical Microbiology. 2014;52(7):2311-2319. DOI: 10.1128/JCM.00598-14. PubMed PMID: 24740082
  97. 97. Haserick JR, Klein JA, Costello CE, Samuelson J. Cryptosporidium parvum vaccine candidates are incompletely modified with O-linked-N-acetylgalactosamine or contain N-terminal N-myristate and S-palmitate. PLoS One. 2017;12(8):e0182395. DOI: 10.1371/journal.pone.0182395. PubMed PMID: 28792526
  98. 98. FDA. New drug for parasitic infections in children. FDA Consumer. 2003;37(3):4. PubMed PMID: 12793375.
  99. 99. Gargala G. Drug treatment and novel drug target against Cryptosporidium. Parasite. 2008;15(3):275-281. DOI: 10.1051/parasite/2008153275. PubMed PMID: 18814694
  100. 100. Singh N, Narayan S. Nitazoxanide: A broad spectrum antimicrobial. Medical Journal, Armed Forces India. 2011;67(1):67-68. DOI: 10.1016/S0377-1237(11)80020- 1S0377-1237(11)80020-1. PubMed PMID: 27365765
  101. 101. Debnath A, Ndao M, Reed SL. Reprofiled drug targets ancient protozoans: Drug discovery for parasitic diarrheal diseases. Gut Microbes. 2013;4(1):66-71. DOI: 10.4161/gmic.22596. PubMed PMID: 23137963
  102. 102. Bessoff K, Sateriale A, Lee KK, Huston CD. Drug repurposing screen reveals FDA-approved inhibitors of human HMG-CoA reductase and isoprenoid synthesis that block Cryptosporidium parvum growth. Antimicrobial Agents and Chemotherapy. 2013;57(4):1804-1814. DOI: 10.1128/AAC.02460-12. PubMed PMID: 23380723
  103. 103. Debnath A, Parsonage D, Andrade RM, He C, Cobo ER, Hirata K, et al. A high-throughput drug screen for Entamoeba histolytica identifies a new lead and target. Nature Medicine. 2012;18(6):956-960. DOI: 10.1038/nm.2758. PubMed PMID: 22610278
  104. 104. Bouzid M, Hunter PR, Chalmers RM, Tyler KM. Cryptosporidium pathogenicity and virulence. Clinical Microbiology Reviews. 2013;26(1):115-134. DOI: 10.1128/CMR.00076-12. PubMed PMID: 23297262
  105. 105. Tzipori S, Griffiths JK. Natural history and biology of Cryptosporidium parvum. Advances in Parasitology. 1998;40:5-36. PubMed PMID: 9554069
  106. 106. Leitch GJ, He Q. Cryptosporidiosis—An overview. Journal of Biomedical Research. 2011;25(1):1-16. DOI: 10.1016/S1674-8301(11)60001-8. PubMed PMID: 22685452
  107. 107. O’Hara SP, Chen XM. The cell biology of Cryptosporidium infection. Microbes and Infection. 2011;13(8-9):721-730. DOI: 10.1016/j.micinf.2011.03.008. PubMed PMID: 21458585
  108. 108. Koh W, Thompson A, Edwards H, Monis P, Clode PL. Extracellular excystation and development of Cryptosporidium: Tracing the fate of oocysts within Pseudomonas aquatic biofilm systems. BMC Microbiology. 2014;14:281. DOI: 10.1186/s12866-014-0281-8. PubMed PMID: 25403949
  109. 109. Aldeyarbi HM, Karanis P. Electron microscopic observation of the early stages of Cryptosporidium parvum asexual multiplication and development in in vitro axenic culture. European Journal of Protistology. 2016;52:36-44. DOI: 10.1016/j.ejop.2015.07.002. PubMed PMID: 26587578
  110. 110. Karanis P, Aldeyarbi HM. Evolution of Cryptosporidium in vitro culture. International Journal for Parasitology. 2011;41(12):1231-1242. DOI: 10.1016/j.ijpara.2011.08.001. PubMed PMID: 21889507
  111. 111. Ryan U, Paparini A, Monis P, Hijjawi N. It’s official-Cryptosporidium is a gregarine: What are the implications for the water industry? Water Research. 2016;105:305-313. DOI: 10.1016/j.watres.2016.09.013. PubMed PMID: 27639055
  112. 112. Tzipori S, Widmer G. A hundred-year retrospective on cryptosporidiosis. Trends in Parasitology. 2008;24(4):184-189. DOI: 10.1016/j.pt.2008.01.002. PubMed PMID: 18329342
  113. 113. Petry F. Structural analysis of Cryptosporidium parvum. Microscopy and Microanalysis. 2004;10(5):586-601. DOI: 10.1017/S1431927604040929. PubMed PMID: 15525433
  114. 114. Liu S, Roellig DM, Guo Y, Li N, Frace MA, Tang K, et al. Evolution of mitosome metabolism and invasion-related proteins in Cryptosporidium. BMC Genomics. 2016;17(1):1006. DOI: 10.1186/s12864-016-3343-5. PubMed PMID: 27931183
  115. 115. Cavalier-Smith T. Gregarine site-heterogeneous 18S rDNA trees, revision of gregarine higher classification, and the evolutionary diversification of Sporozoa. European Journal of Protistology. 2014;50(5):472-495. DOI: 10.1016/j.ejop.2014.07.002. PubMed PMID: 25238406
  116. 116. Johnson AM, Linden K, Ciociola KM, De Leon R, Widmer G, Rochelle PA. UV inactivation of Cryptosporidium hominis as measured in cell culture. Applied and Environmental Microbiology. 2005;71(5):2800-2802. DOI: 10.1128/AEM.71.5.2800-2802.2005. PubMed PMID: 15870378
  117. 117. Robertson LJ, Campbell AT, Smith HV. Survival of Cryptosporidium parvum oocysts under various environmental pressures. Applied and Environmental Microbiology. 1992;58(11):3494-3500. PubMed PMID: 1482175
  118. 118. Deng MQ , Cliver DO. Cryptosporidium parvum studies with dairy products. International Journal of Food Microbiology. 1999;46(2):113-121. PubMed PMID: 10728612
  119. 119. Davies CM, Altavilla N, Krogh M, Ferguson CM, Deere DA, Ashbolt NJ. Environmental inactivation of Cryptosporidium oocysts in catchment soils. Journal of Applied Microbiology. 2005;98(2):308-317. DOI: 10.1111/j.1365-2672.2004.02459.x. PubMed PMID: 15659185
  120. 120. King BJ, Monis PT. Critical processes affecting Cryptosporidium oocyst survival in the environment. Parasitology. 2007;134(Pt 3):309-323. DOI: 10.1017/S0031182006001491. PubMed PMID: 17096874
  121. 121. Jenkins MB, Bowman DD, Fogarty EA, Ghiorse WC. Cryptosporidium parvum oocyst inactivation in three soil types at various temperatures and water potentials. Soil Biology and Biochemistry. 2002;34:1101-1109
  122. 122. Fayer R, Nerad T. Effects of low temperatures on viability of Cryptosporidium parvum oocysts. Applied and Environmental Microbiology. 1996;62(4):1431-1433. PubMed PMID: 8919806
  123. 123. Keevil CW. Rapid detection of biofilms and adherent pathogens using scanning confocal laser microscopy and episcopic differential interference contrast microscopy. Water Science and Technology. 2003;47(5):105-116. PubMed PMID: 12701914
  124. 124. Fayer R, Graczyk TK, Lewis EJ, Trout JM, Farley CA. Survival of infectious Cryptosporidium parvum oocysts in seawater and eastern oysters (Crassostrea virginica) in the Chesapeake Bay. Applied and Environmental Microbiology. 1998;64(3):1070-1074. PubMed PMID: 9501446
  125. 125. World Health Organization. Water, Sanitation and Health Team. Risk Assessment of Cryptosporidium in Drinking Water. Geneva: World Health Organization; 2009
  126. 126. Putignani L, Menichella D. Global distribution, public health and clinical impact of the protozoan pathogen Cryptosporidium. Interdisciplinary Perspectives on Infectious Diseases. 2010;2010:1-39. DOI: 10.1155/2010/753512. PubMed PMID: 20706669
  127. 127. Parsons MB, Travis D, Lonsdorf EV, Lipende I, Roellig DM, Collins A, et al. Epidemiology and molecular characterization of Cryptosporidium spp. in humans, wild primates, and domesticated animals in the greater Gombe ecosystem, Tanzania. PLoS Neglected Tropical Diseases. 2015;9(2):e0003529. DOI: 10.1371/journal.pntd.0003529. PubMed PMID: 25700265
  128. 128. Caccio SM, Sannella AR, Mariano V, Valentini S, Berti F, Tosini F, et al. A rare Cryptosporidium parvum genotype associated with infection of lambs and zoonotic transmission in Italy. Veterinary Parasitology. 2013;191(1-2):128-131. DOI: 10.1016/j.vetpar.2012.08.010. PubMed PMID: 22954678
  129. 129. Casemore DP. Sheep as a source of human cryptosporidiosis. The Journal of Infection. 1989;19(2):101-104. PubMed PMID: 2809233
  130. 130. Current WL, Reese NC, Ernst JV, Bailey WS, Heyman MB, Weinstein WM. Human cryptosporidiosis in immunocompetent and immunodeficient persons. Studies of an outbreak and experimental transmission. The New England Journal of Medicine. 1983;308(21):1252-1257. DOI: 10.1056/NEJM198305263082102. PubMed PMID: 6843609
  131. 131. Efstratiou A, Ongerth JE, Karanis P. Waterborne transmission of protozoan parasites: Review of worldwide outbreaks—An update 2011-2016. Water Research. 2017;114:14-22
  132. 132. Ryan U, Hijjawi N, Xiao L. Foodborne cryptosporidiosis. International Journal for Parasitology. 2018;48(1):1-12. DOI: 10.1016/j.ijpara.2017.09.004. PubMed PMID: 29122606
  133. 133. Hlavsa MC, Roberts VA, Anderson AR, Hill VR, Kahler AM, Orr M, et al. Surveillance for waterborne disease outbreaks and other health events associated with recreational water—United States, 2007-2008. MMWR Surveillance Summaries. 2011;60(12):1-32. PubMed PMID: 21937976
  134. 134. Hunter PR, Zmirou-Navier D, Hartemann P. Estimating the impact on health of poor reliability of drinking water interventions in developing countries. Science of the Total Environment. 2009;407(8):2621-2624. DOI: 10.1016/j.scitotenv.2009.01.018. PubMed PMID: 19193396
  135. 135. Reinoso R, Becares E, Smith HV. Effect of various environmental factors on the viability of Cryptosporidium parvum oocysts. Journal of Applied Microbiology. 2008;104(4):980-986. DOI: 10.1111/j.1365-2672.2007.03620.x. PubMed PMID: 17973913
  136. 136. Chappell CL, Okhuysen PC, Sterling CR, DuPont HL. Cryptosporidium parvum: Intensity of infection and oocyst excretion patterns in healthy volunteers. The Journal of Infectious Diseases. 1996;173(1):232-236. DOI: 10.1093/infdis/173.1.232. PubMed PMID: 8537664
  137. 137. Chappell CL, Okhuysen PC, Sterling CR, Wang C, Jakubowski W, Dupont HL. Infectivity of Cryptosporidium parvum in healthy adults with pre-existing anti-C. parvum serum immunoglobulin G. The American Journal of Tropical Medicine and Hygiene. 1999;60(1):157-164. DOI: 10.4269/ajtmh.1999.60.157. PubMed PMID: 9988341
  138. 138. MacKenzie WR, Schell WL, Blair KA, Addiss DG, Peterson DE, Hoxie NJ, et al. Massive outbreak of waterborne Cryptosporidium infection in Milwaukee, Wisconsin: Recurrence of illness and risk of secondary transmission. Clinical Infectious Diseases. 1995;21(1):57-62. DOI: 10.1093/clinids/21.1.57. PubMed PMID: 7578760
  139. 139. Leoni F, Amar C, Nichols G, Pedraza-Diaz S, McLauchlin J. Genetic analysis of Cryptosporidium from 2414 humans with diarrhoea in England between 1985 and 2000. Journal of Medical Microbiology. 2006;55(Pt 6):703-707. DOI: 10.1099/jmm.0.46251-0. PubMed PMID: 16687587
  140. 140. Morse TD, Nichols RA, Grimason AM, Campbell BM, Tembo KC, Smith HV. Incidence of cryptosporidiosis species in paediatric patients in Malawi. Epidemiology and Infection. 2007;135(8):1307-1315. DOI: 10.1017/S0950268806007758. PubMed PMID: 17224087
  141. 141. Waldron LS, Cheung-Kwok-Sang C, Power ML. Wildlife-associated Cryptosporidium fayeri in human, Australia. Emerging Infectious Diseases. 2010;16(12):2006-2007. DOI: 10.3201/eid1612.100715. PubMed PMID: 21122247
  142. 142. Agholi M, Hatam GR, Motazedian MH. HIV/AIDS-associated opportunistic protozoal diarrhea. AIDS Research and Human Retroviruses. 2013;29(1):35-41. DOI: 10.1089/AID.2012.0119. PubMed PMID: 22873400
  143. 143. Jiang Y, Ren J, Yuan Z, Liu A, Zhao H, Liu H, et al. Cryptosporidium andersoni as a novel predominant Cryptosporidium species in outpatients with diarrhea in Jiangsu Province, China. BMC Infectious Diseases. 2014;14:555. DOI: 10.1186/s12879-014-0555-7. PubMed PMID: 25344387
  144. 144. Liu H, Shen Y, Yin J, Yuan Z, Jiang Y, Xu Y, et al. Prevalence and genetic characterization of Cryptosporidium, Enterocytozoon, Giardia and Cyclospora in diarrheal outpatients in China. BMC Infectious Diseases. 2014;14:25. DOI: 10.1186/1471-2334-14-25. PubMed PMID: 24410985
  145. 145. Ditrich O, Palkovic L, Sterba J, Prokopic J, Loudova J, Giboda M. The first finding of Cryptosporidium baileyi in man. Parasitology Research. 1991;77(1):44-47. PubMed PMID: 1825238
  146. 146. Khan SM, Debnath C, Pramanik AK, Xiao L, Nozaki T, Ganguly S. Molecular characterization and assessment of zoonotic transmission of Cryptosporidium from dairy cattle in West Bengal, India. Veterinary Parasitology. 2010;171(1-2):41-47. DOI: 10.1016/j.vetpar.2010.03.008. PubMed PMID: 20356678
  147. 147. Helmy YA, Krucken J, Nockler K, von Samson-Himmelstjerna G, Zessin KH. Molecular epidemiology of Cryptosporidium in livestock animals and humans in the Ismailia province of Egypt. Veterinary Parasitology. 2013;193(1-3):15-24. DOI: 10.1016/j.vetpar.2012.12.015. PubMed PMID: 23305974
  148. 148. Gatei W, Barrett D, Lindo JF, Eldemire-Shearer D, Cama V, Xiao L. Unique Cryptosporidium population in HIV-infected persons, Jamaica. Emerging Infectious Diseases. 2008;14(5):841-843. DOI: 10.3201/eid1405.071277. PubMed PMID: 18439378
  149. 149. Gatei W, Suputtamongkol Y, Waywa D, Ashford RW, Bailey JW, Greensill J, et al. Zoonotic species of Cryptosporidium are as prevalent as the anthroponotic in HIV-infected patients in Thailand. Annals of Tropical Medicine and Parasitology. 2002;96(8):797-802. DOI: 10.1179/000349802125002202. PubMed PMID: 12625934
  150. 150. Lucio-Forster A, Griffiths JK, Cama VA, Xiao L, Bowman DD. Minimal zoonotic risk of cryptosporidiosis from pet dogs and cats. Trends in Parasitology. 2010;26(4):174-179. DOI: 10.1016/j.pt.2010.01.004. PubMed PMID: 20176507
  151. 151. Robinson G, Chalmers RM. The European rabbit (Oryctolagus cuniculus), a source of zoonotic cryptosporidiosis. Zoonoses and Public Health. 2010;57(7-8):e1-e13. DOI: 10.1111/j.1863-2378.2009.01308.x. PubMed PMID: 20042061
  152. 152. Elwin K, Hadfield SJ, Robinson G, Chalmers RM. The epidemiology of sporadic human infections with unusual cryptosporidia detected during routine typing in England and Wales, 2000-2008. Epidemiology and Infection. 2012;140(4):673-683. DOI: 10.1017/S0950268811000860. PubMed PMID: 21733255
  153. 153. Chalmers RM, Robinson G, Elwin K, Hadfield SJ, Xiao L, Ryan U, et al. Cryptosporidium sp. rabbit genotype, a newly identified human pathogen. Emerging Infectious Diseases. 2009;15(5):829-830. DOI: 10.3201/eid1505.081419. PubMed PMID: 19402985
  154. 154. Chalmers RM, Elwin K, Hadfield SJ, Robinson G. Sporadic human cryptosporidiosis caused by Cryptosporidium cuniculus, United Kingdom, 2007-2008. Emerging Infectious Diseases. 2011;17(3):536-538. DOI: 10.3201/eid1703.100410. PubMed PMID: 21392453
  155. 155. Koehler AV, Whipp MJ, Haydon SR, Gasser RB. Cryptosporidium cuniculus—New records in human and kangaroo in Australia. Parasites & Vectors. 2014;7:492. DOI: 10.1186/s13071-014-0492-8. PubMed PMID: 25359081
  156. 156. Ryan U, Power M. Cryptosporidium species in Australian wildlife and domestic animals. Parasitology. 2012;139(13):1673-1688. DOI: 10.1017/S0031182012001151. PubMed PMID: 22906836
  157. 157. Raccurt CP. Worldwide human zoonotic cryptosporidiosis caused by Cryptosporidium felis. Parasite. 2007;14(1):15-20. DOI: 10.1051/parasite/2007141015. PubMed PMID: 17432054
  158. 158. Cieloszyk J, Goni P, Garcia A, Remacha MA, Sanchez E, Clavel A. Two cases of zoonotic cryptosporidiosis in Spain by the unusual species Cryptosporidium ubiquitum and Cryptosporidium felis. Enfermedades Infecciosas y Microbiología Clínica. 2012;30(9):549-551. DOI: 10.1016/j.eimc.2012.04.011. PubMed PMID: 22728073
  159. 159. Silverlas C, Mattsson JG, Insulander M, Lebbad M. Zoonotic transmission of Cryptosporidium meleagridis on an organic Swedish farm. International Journal for Parasitology. 2012;42(11):963-967. DOI: 10.1016/j.ijpara.2012.08.008. PubMed PMID: 23022616
  160. 160. Cama VA, Ross JM, Crawford S, Kawai V, Chavez-Valdez R, Vargas D, et al. Differences in clinical manifestations among Cryptosporidium species and subtypes in HIV-infected persons. The Journal of Infectious Diseases. 2007;196(5):684-691. DOI: 10.1086/519842. PubMed PMID: 17674309
  161. 161. Palmer CJ, Xiao L, Terashima A, Guerra H, Gotuzzo E, Saldias G, et al. Cryptosporidium muris, a rodent pathogen, recovered from a human in Peru. Emerging Infectious Diseases. 2003;9(9):1174-1176. DOI: 10.3201/eid0909.030047. PubMed PMID: 14519260
  162. 162. Al-Brikan FA, Salem HS, Beeching N, Hilal N. Multilocus genetic analysis of Cryptosporidium isolates from Saudi Arabia. Journal of the Egyptian Society of Parasitology. 2008;38(2):645-658. PubMed PMID: 18853635
  163. 163. Muthusamy D, Rao SS, Ramani S, Monica B, Banerjee I, Abraham OC, et al. Multilocus genotyping of Cryptosporidium sp. isolates from human immunodeficiency virus-infected individuals in South India. Journal of Clinical Microbiology. 2006;44(2):632-634. DOI: 10.1128/JCM.44.2.632-634.2006. PubMed PMID: 16455931
  164. 164. Kvac M, Kvetonova D, Sak B, Ditrich O. Cryptosporidium pig genotype II in immunocompetent man. Emerging Infectious Diseases. 2009;15(6):982-983. DOI: 10.3201/eid1506.071621. PubMed PMID: 19523313
  165. 165. Wang L, Zhang H, Zhao X, Zhang L, Zhang G, Guo M, et al. Zoonotic Cryptosporidium species and Enterocytozoon bieneusi genotypes in HIV-positive patients on antiretroviral therapy. Journal of Clinical Microbiology. 2013;51(2):557-563. DOI: 10.1128/JCM.02758-12. PubMed PMID: 23224097
  166. 166. Xiao L, Bern C, Arrowood M, Sulaiman I, Zhou L, Kawai V, et al. Identification of the Cryptosporidium pig genotype in a human patient. The Journal of Infectious Diseases. 2002;185(12):1846-1848. DOI: 10.1086/340841. PubMed PMID: 12085341
  167. 167. Bodager JR, Parsons MB, Wright PC, Rasambainarivo F, Roellig D, Xiao L, et al. Complex epidemiology and zoonotic potential for Cryptosporidium suis in rural Madagascar. Veterinary Parasitology. 2015;207(1-2):140-143. DOI: 10.1016/j.vetpar.2014.11.013. PubMed PMID: 25481280
  168. 168. Raskova V, Kvetonova D, Sak B, McEvoy J, Edwinson A, Stenger B, et al. Human cryptosporidiosis caused by Cryptosporidium tyzzeri and C. parvum isolates presumably transmitted from wild mice. Journal of Clinical Microbiology. 2013;51(1):360-362. DOI: 10.1128/JCM.02346-12. PubMed PMID: 23100342
  169. 169. Li N, Xiao L, Alderisio K, Elwin K, Cebelinski E, Chalmers R, et al. Subtyping Cryptosporidium ubiquitum,a zoonotic pathogen emerging in humans. Emerging Infectious Diseases. 2014;20(2):217-224. DOI: 10.3201/eid2002.121797. PubMed PMID: 24447504
  170. 170. Elwin K, Hadfield SJ, Robinson G, Crouch ND, Chalmers RM.Cryptosporidium viatorum n. sp. (Apicomplexa: Cryptosporidiidae) among travellers returning to Great Britain from the Indian subcontinent, 2007-2011. International Journal for Parasitology. 2012;42(7):675-682. DOI: 10.1016/j.ijpara.2012.04.016. PubMed PMID: 22633952
  171. 171. Garcia RJ, French N, Pita A, Velathanthiri N, Shrestha R, Hayman D. Local and global genetic diversity of protozoan parasites: Spatial distribution of Cryptosporidium and Giardia genotypes. PLoS Neglected Tropical Diseases. 2017;11(7):e0005736. DOI: 10.1371/journal.pntd.0005736. PubMed PMID: 28704362
  172. 172. Kvac M, Hofmannova L, Hlaskova L, Kvetonova D, Vitovec J, McEvoy J, et al. Cryptosporidium erinacei n. sp. (Apicomplexa: Cryptosporidiidae) in hedgehogs. Veterinary Parasitology. 2014;201(1-2):9-17. DOI: 10.1016/j.vetpar.2014.01.014. PubMed PMID: 24529828
  173. 173. Azami M, Moghaddam DD, Salehi R, Salehi M. The identification of Cryptosporidium species (protozoa) in Ifsahan, Iran by PCR-RFLP analysis of the 18S rRNA gene. Molekuliarnaia Biologiia. 2007;41(5):934-939. PubMed PMID: 18240576
  174. 174. Adamu H, Petros B, Zhang G, Kassa H, Amer S, Ye J, et al. Distribution and clinical manifestations of Cryptosporidium species and subtypes in HIV/AIDS patients in Ethiopia. PLoS Neglected Tropical Diseases. 2014;8(4):e2831. DOI: 10.1371/journal.pntd.0002831. PubMed PMID: 24743521
  175. 175. Ng-Hublin JS, Combs B, Mackenzie B, Ryan U. Human cryptosporidiosis diagnosed in Western Australia: A mixed infection with Cryptosporidium meleagridis, the Cryptosporidium mink genotype, and an unknown Cryptosporidium species. Journal of Clinical Microbiology. 2013;51(7):2463-2465. DOI: 10.1128/JCM.00424-13. PubMed PMID: 23637295
  176. 176. Slapeta J. Cryptosporidiosis and Cryptosporidium species in animals and humans: A thirty colour rainbow? International Journal for Parasitology. 2013;43(12-13):957-970. DOI: 10.1016/j.ijpara.2013.07.005. PubMed PMID: 23973380
  177. 177. Pintar KD, Pollari F, Waltner-Toews D, Charron DF, McEwen SA, Fazil A, et al. A modified case-control study of cryptosporidiosis (using non-Cryptosporidium-infected enteric cases as controls) in a community setting. Epidemiology and Infection. 2009;137(12):1789-1799. DOI: 10.1017/S0950268809990197. PubMed PMID: 19527550
  178. 178. Roy SL, DeLong SM, Stenzel SA, Shiferaw B, Roberts JM, Khalakdina A, et al. Risk factors for sporadic cryptosporidiosis among immunocompetent persons in the United States from 1999 to 2001. Journal of Clinical Microbiology. 2004;42(7):2944-2951. DOI: 10.1128/JCM.42.7.2944-2951.2004. PubMed PMID: 15243043
  179. 179. Yoder JS, Beach MJ. Cryptosporidium surveillance and risk factors in the United States. Experimental Parasitology. 2010;124(1):31-39. DOI: 10.1016/j.exppara.2009.09.020. PubMed PMID: 19786022
  180. 180. King P, Tyler KM, Hunter PR. Anthroponotic transmission of Cryptosporidium parvum predominates in countries with poorer sanitation: A systematic review and meta-analysis. Parasites & Vectors. 2019;12(1):16. DOI: 10.1186/s13071-018-3263-0. PubMed PMID: 30621759
  181. 181. Nazemalhosseini-Mojarad E, Feng Y, Xiao L. The importance of subtype analysis of Cryptosporidium spp. in epidemiological investigations of human cryptosporidiosis in Iran and other mideast countries. Gastroenterology and Hepatology from Bed to Bench. 2012;5(2):67-70. PubMed PMID: 24834202
  182. 182. Heiges M, Wang H, Robinson E, Aurrecoechea C, Gao X, Kaluskar N, et al. CryptoDB: A Cryptosporidium bioinformatics resource update. Nucleic Acids Research. 2006;34:D419-D422. DOI: 10.1093/nar/gkj078. PubMed PMID: 16381902
  183. 183. Abrahamsen MS, Templeton TJ, Enomoto S, Abrahante JE, Zhu G, Lancto CA, et al. Complete genome sequence of the apicomplexan, Cryptosporidium parvum. Science. 2004;304(5669):441-445. DOI: 10.1126/science.1094786. PubMed PMID: 15044751
  184. 184. Guo Y, Tang K, Rowe LA, Li N, Roellig DM, Knipe K, et al. Comparative genomic analysis reveals occurrence of genetic recombination in virulent Cryptosporidium hominis subtypes and telomeric gene duplications in Cryptosporidium parvum. BMC Genomics. 2015;16:320. DOI: 10.1186/s12864-015-1517-1. PubMed PMID: 25903370
  185. 185. Feng Y, Li N, Roellig DM, Kelley A, Liu G, Amer S, et al. Comparative genomic analysis of the IId subtype family of Cryptosporidium parvum. International Journal for Parasitology. 2017;47(5):281-290. DOI: 10.1016/j.ijpara.2016.12.002. PubMed PMID: 28192123
  186. 186. Mogi T, Kita K. Diversity in mitochondrial metabolic pathways in parasitic protists Plasmodium and Cryptosporidium. Parasitology International. 2010;59(3):305-312. DOI: 10.1016/j.parint.2010.04.005. PubMed PMID: 20433942
  187. 187. Guy RA, Payment P, Krull UJ, Horgen PA. Real-time PCR for quantification of Giardia and Cryptosporidium in environmental water samples and sewage. Applied and Environmental Microbiology. 2003;69(9):5178-5185. DOI: 10.1128/aem.69.9.5178-5185.2003. PubMed PMID: 12957899

Written By

Helena Lúcia Carneiro Santos, Karina Mastropasqua Rebello and Teresa Cristina Bergamo Bomfim

Submitted: 02 April 2019 Reviewed: 03 July 2019 Published: 10 September 2019