Find information on thousands of medical conditions and prescription drugs.

Mitochondrial Diseases

Mitochondrial diseases are a group of disorders relating to the mitochondria, the organelles that are the "powerhouses" of the eukaryotic cells that comprise higher-order lifeforms (including humans). The mitochondria convert the energy of food molecules into the ATP that powers most cell functions. more...

Home
Diseases
A
B
C
D
E
F
G
H
I
J
K
L
M
Mac Ardle disease
Macroglobulinemia
Macular degeneration
Mad cow disease
Maghazaji syndrome
Mal de debarquement
Malaria
Malignant hyperthermia
Mallory-Weiss syndrome
Malouf syndrome
Mannosidosis
Marburg fever
Marfan syndrome
MASA syndrome
Mast cell disease
Mastigophobia
Mastocytosis
Mastoiditis
MAT deficiency
Maturity onset diabetes...
McArdle disease
McCune-Albright syndrome
Measles
Mediterranean fever
Megaloblastic anemia
MELAS
Meleda Disease
Melioidosis
Melkersson-Rosenthal...
Melophobia
Meniere's disease
Meningioma
Meningitis
Mental retardation
Mercury (element)
Mesothelioma
Metabolic acidosis
Metabolic disorder
Metachondromatosis
Methylmalonic acidemia
Microcephaly
Microphobia
Microphthalmia
Microscopic polyangiitis
Microsporidiosis
Microtia, meatal atresia...
Migraine
Miller-Dieker syndrome
Mitochondrial Diseases
Mitochondrial...
Mitral valve prolapse
Mobius syndrome
MODY syndrome
Moebius syndrome
Molluscum contagiosum
MOMO syndrome
Mondini Dysplasia
Mondor's disease
Monoclonal gammopathy of...
Morquio syndrome
Motor neuron disease
Motorphobia
Moyamoya disease
MPO deficiency
MR
Mucopolysaccharidosis
Mucopolysaccharidosis...
Mullerian agenesis
Multiple chemical...
Multiple endocrine...
Multiple hereditary...
Multiple myeloma
Multiple organ failure
Multiple sclerosis
Multiple system atrophy
Mumps
Muscular dystrophy
Myalgic encephalomyelitis
Myasthenia gravis
Mycetoma
Mycophobia
Mycosis fungoides
Myelitis
Myelodysplasia
Myelodysplastic syndromes
Myelofibrosis
Myeloperoxidase deficiency
Myoadenylate deaminase...
Myocarditis
Myoclonus
Myoglobinuria
Myopathy
Myopia
Myositis
Myositis ossificans
Myxedema
Myxozoa
N
O
P
Q
R
S
T
U
V
W
X
Y
Z
Medicines

Mitochondrial diseases comprise those disorders that in one way or another affect the function of the mitochondria and/or are due to mitochondrial DNA. Mitochondrial diseases take on unique characteristics both because of the way the diseases are often inherited and because that mitochondria are so critical to cell function. The subclass of these diseases that have neuromuscular disease symptoms are often referred to as a mitochondrial myopathy.

Mitochondrial inheritance

Mitochondrial inheritance behaves differently from the sort of inheritance that we are most familiar with. Regular nuclear DNA has two copies per cell (except for sperm and egg cells). One copy is inherited from the father and the other from the mother. Mitochondria, however, contain their own DNA, and contain typically from five to ten copies, all inherited from the mother (for more detailed inheritance patterns, see mitochondrial genetics). When mitochondria divide, the copies of DNA present are divided randomly between the two new mitochondria, and then those new mitochondria make more copies. As a result, if only a few of the DNA copies inherited from the mother are defective, mitochondrial division may cause most of the defective copies to end up in just one of the new mitochondria. Once more than half of the DNA copies are defective, mitochondrial disease begins to become apparent, this phenomenon is called 'threshold expression'.

It should be noted, however, that not all of the enzymes and other components necessary for proper mitochondrial function are encoded in the mitochondrial DNA. Most mitochondrial function is controlled by nuclear DNA instead.

To make things even more confusing, mutations to mitochondrial DNA occur frequently, due to the lack of the error checking capability that nuclear DNA has. This means that a mitochondrial disorder can occur spontaneously rather than be inherited. Further, sometimes the enzymes that control mitochondrial DNA duplication (and which are encoded for by genes in the nuclear DNA) are defective, causing mitochondrial DNA mutations to occur at a rapid rate.

Defects and symptoms

The effects of mitochondrial disease can be quite varied. Since the distribution of defective DNA may vary from organ to organ within the body, the mutation that in one person may cause liver disease might in another person cause a brain disorder. In addition, the severity of the defect may be great or small. Some minor defects cause only "exercise intolerance", with no serious illness or disability. Other defects can more severely affect the operation of the mitochondria and can cause severe body-wide impacts. As a general rule, mitochondrial diseases are worst when the defective mitochondria are present in the muscles or nerves, because these are the most energy-hungry cells of the body.

Read more at Wikipedia.org


[List your site here Free!]


Isolation of mitochondrial cytochrome B gene and development of a real-time quantitative PCR assay for detecting Neoparamoeba aestuarina
From Journal of Shellfisheries Research, 10/1/05 by Senjie Lin

ABSTRACT Some Neoparamoeba strains are pathogens of fish and invertebrates and were implicated in lobster mortality in Long Island Sound in recent years. To better understand the dynamics of these pathogens and their potential linkages to the mortality of marine animals, the capability to specifically detect and accurately quantify these parasites in the environment and in affected organisms is essential. Molecular markers such as mitochondrial cytochrome b (cob) can be powerful tools. In this study, we isolated cob from N. aestuarina and developed species-specific primers. Sequence analysis showed that this gene was A/T rich. Variable regions were identified and used to develop a species-specific primer set for real-time quantitative PCR (RTQ-PCR). This primer set was demonstrated to be specific (with no cross-reaction to N. pemaquidensis and other protists) and sensitive (with a detection limit of 1 cell/30 mL). Further, the quantitative capability of the RTQ-PCR system was verified by a strong correlation between the threshold cycle number and the logarithmic-transformed cob copy number used in the amplification. Using the established RTQ-PCR assay, we estimated that each N. aestuarina cell contained approximately 707 [+ or -] 63 copies of cob. This assay was applied to Lugol-fixed water samples collected from Long Island Sound during August 2002 to September 2003, which showed no detectable N. aestuarina. The cob primer set developed will be useful for future environmental detection of N. aestuarina and provides a basis from which cob gene probes for other Neoparamoeba and Paramoeba species can be developed.

KEY WORDS: cob, mitochondrial cytochrome b, Neoparamoeba aestuarina, Real-Time quantitative PCR (RTQ-PCR), species-specific gene probe

INTRODUCTION

Paramoebid protists (Sarcomastigophora: Paramoebidae) are common marine protozoa (Page 1973), some of which, especially Paramoeba Schaudinn, 1896 and Neoparamoeba Page, 1987, have been linked to diseases and death in fish and invertebrates. Paramoeba perniciosa was first identified in association with "gray crab" disease in blue crabs (Callinectes sapidus) in Maryland, Virginia, North Carolina and South Carolina (Sprague & Beckett 1966, Sprague & Beckett 1968, Sprague et al. 1969, Newman & Ward 1973, Johnson 1977). In blue crabs, P. perniciosa is localized in hemal spaces and connective tissues and spreads to blood vessels and eventually to the heart during the terminal stage of infection (Johnson 1977). This parasite also invades muscle tissue and rarely, hepatopancreas, and in heavy infection can be found in central nervous system. Paramoeba invadens has been linked to massive mortality of sea urchin (Strongylocentrotus droebachiensis) in Nova Scotia, in which the parasite is widespread in all tissues but it is consistently localized in the radial nerve/water vascular canal tissue (Jones & Scheibling 1985, Jellett et al. 1988). More recently, N. pemaquidensis was also recognized as a pathogen of salmon and rainbow trout (Kent et al. 1988). In contrast to the Paramoeba species as mentioned earlier, N. pemaquidensis mostly infects the gills of salmon and other fish (Kent et al. 1988, Dykova et al. 2000). N. aestuarina has also been suggested as a potential agent of amoebic gill disease (Dykova et al. 2000). In addition, Neoparamoeba spp. have been shown to raise immune response in fish (Gross et al. 2004).

Protozoa in the Neoparamoeba/Paramoeba complex of species have been implicated in the massive mortality of the American lobster (Homarus americanus) in Long Island Sound in the fall of 1999 (CT DEP 1999, Tedesco & Van Patten 2000). In diseased and dead lobsters, paramoebid parasites have been identified most frequently in ganglia of the central nervous system (Russell et al. 2000). Although it remains unclear whether paramoebiasis is the primary cause of the die-off, it is believed to be at least involved in some stages of the sickening process (e.g., Russell et al. 2000, Mullen et al. 2004). Furthermore, paramoebid parasites may act synergistically with other stressors such as elevated temperature (Scheibling & Hennigar 1997), which occurred in Long Island Sound bottom water in 1999 (R. Wilson, pers. comm.). To correlate the shellfish disease and lobster mortality with Neoparamoeba/Paramoeba, a capability to detect and quantify these pathogens specifically and efficiently is essential.

Identification of Neoparamoeba/Paramoeba is challenging. Although morphologic and histologic characteristics may allow separation of the documented species (e.g., Perkins & Castagna 1971), it is not practical in environmental or epidemiologic studies to process large number of samples for those characteristics. Molecular markers have proven to be powerful tools for specifically detecting and accurately quantifying microorganisms on large sample scales. Among potentially useful genes, mitochondrial (mt) genes are good candidates because they are generally conserved and yet their higher variability than some nuclear genes (Arise 1994, Saccone et al. 2000) renders these genes preferable for development of species-specific probes. Mt genes also appear to be more clock-like (i.e., of relatively even mutation rate across taxa), characteristics essential for unbiased phylogenetic separation of different taxa (Saccone et al. 2000). Cytochrome b (cob) is one of the mt genes that is widely used for phylogenetic and population genetic analyses (e.g., Callejas & Ochando 2000, Taylor & Hellberg 2003, Zhang et al. 2005). Employment of cob for detecting harmful marine protists has just begun (e.g., Zhang & Lin 2002) and has not been documented for amoeboid species.

In this study, we attempt to isolate cob from N. aestuarina and N. pemaquidensis and developed specific probes for each species. They were chosen because paramoebid parasites detected from infected Long Island Sound lobsters were most closely related to N. pemaquidensis (Mullen et al. 2005) and N. aestuarina is phylogenetically close to N. pemaquidensis (Peglar et al. 2003, Wong et al. 2004). We had success only with N. aestuarina and hence developed a real-time quantitative PCR (RTQ-PCR) assay only for this species at this time.

MATERIALS AND METHODS

Neoparamoeba Cultures

Neoparamoeba aestuarina (ATCC 50744) and N. pemaquidensis (ATCC 30735) were provided by American Type Culture Center (ATCC, Manassas, VA) as nonaxenic cultures fed with polyclonal bacteria dominated by Klebsiella pneumoniae (ATCC 700831). Once the cultures, shipped on dry ice, arrived in the laboratory, cells were thawed, transferred to 1.5 mL microcentrifuge tubes, and harvested by centrifugation at x3000g at 4[degrees]C for 5 min. The cell density in the cultures was approximately 1 million cells/mL

DNA Extraction

DNA was extracted according to Zhang and Lin (2002). Briefly, cells were harvested by centrifugation at x3000g for 20 min at 4[degrees]C. Cell pellets were resuspended in 400 [micro]L DNA extraction buffer (0.1 M EDTA, 1% SDS, 0.2 mg/mL Proteinase K) and incubated overnight at 55[degrees]C. The crude DNA solutions were purified by phenol/chloroform (1:1 vol/vol) extraction. DNA was then precipitated by adding 20 [micro]L of 3M sodium acetate (pH 7) and 800 [micro]L of 99.9% ethyl ethanol, incubating at -20[degrees]C for 1 h. After centrifugation at x16000g for 10 min at 4[degrees]C, DNA pellets were washed with 75% ethanol, dried and dissolved in 10 mM Tris-HCl (pH 8.0) at a concentration equivalent to 4 x [10.sup.5] cells/ mL. Difficulties were encountered in separation of the N. pemaquidensis cells from agar, some of which coprecipitated with DNA. Therefore, more extensive phenol extraction was carried out to achieve DNA that was PCR-amplifiable as demonstrated with 18S rDNA primers.

Primer Design and PCR

Lack of cob sequence information for amoeboid organisms made the primer design a challenge. Primers recently designed and demonstrated to be effective for a wide range of dinoflagellates (Zhang et al. 2005) were first tested but did not generate results. Therefore, new primers were designed based on a multi-alignment including several dinoflagellates and Acanthamoeba castellanii (strain ATCC 30010), the only amoebid protozoa in which a cob sequence (derived from complete mt DNA sequence) is available in GenBank. Two highly conserved regions were identified from the amino acid sequence alignment (Fig. 1). Six "Universal" degenerate primers (Table 1) were designed using consensusdegenerate hybrid oligonucleotide primers (CODEHOP) software (http://blocks.fhcrc.org/codehop.html). Combinations of these primers with dinoflagellate primers were used in PCR with genomic DNA as the template. Temperature-gradient PCR was carried with a PCR express thermocycler (Hybaid, Action Court, Middlesex, UK) with one single cycle of 3 min at 94[degrees]C followed by 35 cycles of 94[degrees]C for 30 sec, 42[degrees]C to 58[degrees]C for 30 sec (with 1[degrees]C increments), 72[degrees]C for 30 sec, and ended with one additional step of 72[degrees]C for 7 min. Results were analyzed on agarose gel electrophoresis.

[FIGURE 1 OMITTED]

Cloning and Sequencing

PCR products were excised from agarose gel, purified using Zymoclean Gel DNA Recovery Kit (Zymo Research, Orange, CA), and then ligated to a TA vector (Invitrogen). The recombinant DNA was transformed into XL1-Blue E. coli strain. Twenty clones were randomly picked for each PCR product and subjected to sequencing on ABI Prism 377 automated sequencer using Big-Dye Terminator Cycle Sequencing Kit (PE Applied Biosystems, Foster City, CA).

Phylogenetic Analysis

The nucleotide sequences were aligned using the CLUSTAL W (1.8) server at DDBJ (http://www.ddbj.nig.ac.jp/Welcome-e.html) on default values, and adjusted using the SeaView program to maintain codon integrity prior to phylogenetic analysis in the Phylo_Win package (Galtier et al. 1996). Phylogenetic trees were constructed from the nt data by the neighbor-joining method (Saitou & Nei 1987) and corrected using Kimura 2 parameter model (Kimura 1980).

Development of Species-specific Primers and Real-Time PCR

N. aestuarina cob sequences were aligned with orthologs from other unicellular organisms. Regions unique in N. aestuarina were identified and used to design species-specific primers using Beacon Designer version 2.1 software (Premier Biosoft International, Palo Alto, CA). Real-Time PCR was carried out with SYBR Green Supermix on an iCycler iQ Real-Time PCR Detection System (Bio-Rad Laboratories, Inc., Hercules, CA). Amplification conditions were 45 cycles of 20 sec of denaturation at 95[degrees]C, 30 sec of annealing at 58[degrees]C, and 20 sec of elongation at 72[degrees]C. Data on DNA amount were acquired at the end of extension step in each PCR cycle. The software that controlled the system analyzed the results automatically.

Field Sample Analysis

Samples were collected from four stations (A4, 09, H4 and K2) in Long Island Sound along a west-east axis (Fig. 2). The lobster mortality was most severe in the western Sound, where episodic hypoxia has recurred. Water samples were taken, mostly monthly-bimonthly (Table 2), onboard R/V Dempsey at 2 m depth with Niskin bottles mounted on a CTD rosette. Subsamples of 200 mL were taken and preserved immediately with Lugol iodine solution (Utermohl 1958) and transported to our laboratory within a week. Samples were then stored at 4[degrees]C until analysis. For each sample, 30 mL was removed from the sample bottle, and cells were concentrated by centrifugation at x3000g in two 15-mL tubes with a swing bucket for 20 min at 4[degrees]C. The cell pellet was subjected to DNA extraction as described earlier and dissolved in 30-[micro]L autoclaved distilled water. From each sample, 1 [micro]L of the 30 [micro]L DNA solution was used in Real-Time PCR using the specific primers.

[FIGURE 2 OMITTED]

RESULTS AND DISCUSSION

cob Gene in Neoparamoeba

PCR amplification of cob in these Neoparamoeba species proved difficult. Initial attempts using all possible combinations of the available primers (Table 1) at 55[degrees]C annealing temperature did not yield any positive results. Use of a lower annealing temperature (48[degrees]C) with primers cob5d and cob3a (Fig. 1, Table 1) generated a product of 433 bp for N. aestuarina (GenBank accession number AY743963) and a product of 436 for N. pemaquidensis (GenBank accession number AY743964). Sequencing of the cloned products followed by BLAST search showed that the former was a cob sequence closely related to eukaryotic homologs, but the latter was most similar to sequences of cytochrome bc1 complex genes in [alpha]-proteobacbacteria.

Phylogenetic analyses including the sequences yielded from the Neoparamoeba strains, Acanthamoeba, the only close relative of Neoparamoeba in which cob sequence was available, and some other eukaryotic and prokaryotic unicellular organisms showed that the N. aestuarina-derived cob clustered with eukaryotes and the N. pemaquidensis-derived sequence with proteobacteria (Fig. 3). These phylogenetic placements, with strong bootstrap support, reinforced the BLAST analysis result shown earlier. The PCR product for N. pemaquidensis was most likely derived from bacteria present in the N. pemaquidensis culture (e.g., Klebsiella pneumoniae [ATCC 700831] provided as food). The failure to isolate cob from N. pemaquidensis suggests difference in cob sequence between N. aestuarina and N. pemaquidensis (e.g., variation in intron structure). Hence, only the N. aestuarina-derived cob sequence was analyzed further.

[FIGURE 3 OMITTED]

As Neoparamoeba spp. are commonly associated endocytotically with Perkinsella sp. (Kinetoplastida), the so-called parasome (Dykova et al. 2003), the N. aestuarina-derived cob can theoretically be Neoparamoeba- or parasome-origin. However, it is highly unlikely that the primers used would have amplified cob of the parasome, because kinetoplastida cob genes are highly "cryptic" with numerous deletions or insertions of thymidine (Scott 1995, Stuart et al. 2001). Even if the parasome cob gene is amplifiable, its sequence would not likely be recognizable without addition or removal of thymidine. Therefore, the cloned cob gene fragment is considered a genuine mitochondrial cob from N. aestuarina.

Species-specific Primers

When N. aestuarina cob sequence was aligned with homologs from other organisms, including Acanthamoeba, highly variable regions of the gene were identifiable (Fig. 4). Using Beacon Designer version 2.1, with G/C content and self-annealing taken into consideration, a pair of primers, ParacobF and ParacobR, were obtained (Table 1).

[FIGURE 4 OMITTED]

For the primer set to be useful for field studies, specificity, quantification capability and sensitivity need to be demonstrated. Specificity was first tested with DNA extracted from IV. aestuarina and N. pemaquidensis. Using annealing temperatures 50[degrees]C to 58[degrees]C, PCR with the primer set did not generate any product for N. pemaquidensis, whereas a single DNA product was consistently yielded from N. aestuarina (Fig. 5). No amplicon was obtained from other protist cultures such as dinoflagellates (not shown). The specificity of the primer set was also verified by the lack of PCR product from the DNA extracted from water samples collected from eastern Long Island Sound at the Avery Point campus of University of Connecticut, in which the plankton assemblage consists of numerous protist species.

[FIGURE 5 OMITTED]

Real-Time PCR was used to determine whether this primer set would allow quantification of N. aestuarina with a reasonable detection limit. First, a dilution series of the cloned cob plasmid was used in Real-Time PCR to construct a standard curve (Fig. 6). The amplification product increased with gene copy number (Fig. 5), and, therefore, threshold cycle number decreased exponentially with gene copy number (Fig. 6). A very good correlation was attained between the threshold cycle number and the logarithmic-transformed gene copy number (Fig. 6); thus, the quantitative capability of the Real-Time PCR system is demonstrated. Further analysis indicated that the system was able to detect cob plasmid at as low a dose as 10 copies.

[FIGURE 6 OMITTED]

Second, the RTQ-PCR system was used to estimate cob copy number in N. aestuarina and then to recalculate detection limits in terms of the number of N. aestuarina cells. A dilution series of N. aestuarina DNA was subjected to Real-Time PCR as for the dilution series of cob plasmid described earlier. Proportional increases in PCR product (or decrease of threshold cycle number) with increases of DNA concentration was observed (Fig. 6). Using the standard curve constructed with the plasmid dilution series, the total number of cob copies from each DNA dilution was derived. By averaging the results from all DNA dilutions (n = 8), an estimate of 707 [+ or -] 63 copies per cell was obtained. Finally, based on this estimate, the theoretical detection limit was estimated as 10/707 = 0.014 cell/reaction. Because we normally isolate DNA from 30-mL field-collected water samples, dissolve DNA in 30 [micro]L autoclaved distilled water and then use 1 [micro]L for Real-Time PCR reaction, the practical detection limit of this system is 1 cell per 30 mL of water sample (or 0.03 cell/mL). Therefore, this PCR based cob detection system is in theory considerably more sensitive than whole-cell staining methods (e.g., Lin et al. 2003).

Detection of N. aestuarina in Long Island Sound

Using the RTQ-PCR system, no positive results were generated from any of the DNA samples extracted from water samples collected along the west-east axis of Long Island Sound during August 2002 to September 2003 (Fig. 2, Table 2). The same DNA samples have been used recently to analyze diversity of dinoflagellates and small-sized phytoplankton communities and shown to be of PCR-amplification quality (Zhang & Lin unpublished data). To further examine whether the negative results were because the primer set or PCR conditions used in this study were more susceptible to inhibitory chemicals that might exist in the samples, N. aestuarina DNA extracted from the culture (standard DNA) was spiked with DNA isolated from Long Island Sound water samples and amplified in parallel with the pure standard DNA. No difference was found between the spiked and nonspiked DNA. All these results indicate that the negative results from the field samples were not due to failure of PCR, rather, N. aestuarina was not present above, if at all, the detection limit in the samples. There is a possibility that the absence of N. aestuarina in the samples was a result of sample fixation and storage in Lugol solution. This fixative has been used widely to preserve plankton cells and extensive tests with dinoflagellates have shown that it preserves cells effectively (Zhang & Lin 2002 and unpubl.), but has not been examined for N. aestuarina. If Lugol solution proves effective in preserving N. aestuarina then this negative result would suggest that N. aestuarina is not likely one of the diverse and widespread Neoparamoeba species that have recently been detected in Long Island Sound using SSU rDNA (Gast 2004). It would follow that whereas the SSU rDNA-based analysis has shown that the paramoebid parasite infecting the diseased lobster from Long Island Sound is closely related to N. pemaquidensis (Mullen et al. 2004), it is not likely to be N. aestuarina.

CONCLUSION

This study demonstrates that (1) cob is highly A/T rich and challenging to clone in Neoparamoeba (as in protistan cob genes in general); (2) each N. aestuarina cell contains approximately 707 [+ or -] 63 copies of cob; (3) the cob-based RTQ-PCR system developed is useful for detecting and quantifying N. aestuarina.

Further research is required to clone cob from other Neoparamoeba species to establish capability to detect and quantify other paramoebid pathogens. Once more species are investigated, dualgene primers like cob-SSU rDNA primer set developed for Priesteria piscicida (Zhang & Lin 2002) may prove desirable for distinguishing some of the closely related species. These multiple genes would also be very useful for phylogenetically resolving these taxa.

ACKNOWLEDGMENTS

The authors thank Connecticut Sea Grant for development funds to support this study under Grant No. LMP/D-1, through the US Department of Commerce, National Oceanic and Atmospheric Administration (NOAA) Award NA16RGI364, and the Water Quality Monitoring group of the Connecticut Department of Environmental Protection for the water samples.

LITERATURE CITED

Avise, J. C. 1994. Molecular markers, natural history and evolution, New York: Chapman & Hall. 511 pp.

Callejas, C. & M. D. Ochando. 2000. Recent radiation of Iberian barbel fish (Teleostei, Cyprinidae) inferred from cytochrome b genes. J. Hered. 91:283-288.

Connecticut Department of Environmental Protection. 1999. Information regarding the impact of 1999 lobster mortalities in Long Island Sound. CT DEP Marine Fisheries Office, Old Lyme, CT. 45 pp.

Dykova, I., A. Figueras & Z. Peric. 2000. Neoparamoeba Page, 1987: light and electron microscopic observations on six strains of different origin. Dis. Aquat. Organ. 43:217-223.

Dykova, I., I. Fiala, J. Lore & J. Lukes. 2003. Perkinsiella amoebae-like endosymbionts of Neoparamoeba spp., relatives of the kinetoplastid Ichthyobodo. Eur. J. Protistol. 39:37-52.

Galtier, N., M. Gouy & C. Gautier. 1996. SEAVIEW and PHYLO_WIN: two graphic tools for sequence alignment and molecular phylogeny. Computer Applications in the Biosciences 12:543-548.

Gast, R. J. 2004. Oligonucleotide-based detection of pathogenic Neoparamoeba species. Abstract, second working meeting of LIS Lobster Research Initiative. May 3-4. Groton, Connecticut.

Gross, K., J. Carson & B. Nowak. 2004. Presence of anti-Neoparamoeba sp. antibodies in Tasmanian cultured Atlantic salmon, Salmo salar L. J. Fish Dis. 27:81-88.

Jellett, J. F., A. C. Wardlaw & R. E. Scheibling. 1988. Experimental infection of the echinoid Strongylocentrotus droebachiensis with Paramoeba invadens: quantitative changes in the coelomic fluid. Dis. Aquat. Org. 4:149-157.

Johnson, P. T. 1977. Paramoebiasis in the blue crab, Callinectes sapidus. J. Invert. Path. 29:308-320.

Jones, G. M. & R. E. Scheibling. 1985. Paramoeba sp. (Amoebida, Paramoebidae) as the possible causative agent of sea urchin mass mortality in Nova Scotia. J. Parasitol. 71:559-565.

Kent, M. L., T. K. Sawyer & R. P. Hedrick. 1988. Paramoeba pemaquidensis (Sarcomastigophora: Paramoebidae) infestation of the gills of coho salmon Oncorhynchus kisutch reared in sea water. Dis. Aquat. Org. 5:163-169.

Kimura, M. 1980. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 16:111-120.

Lin, S., T. N. Feinstein, H. Zhang & E. J. Carpenter. 2003. Development of an immuno-fluorescence technique for detecting Pfiesteria piscicida. Harmful Algae 2:223-231.

Mullen, T. E., S. Russell, M. T. Tucker, J. L. Maratea, T. G. Burrage, C. Koerting, L. Hinckley, C. R. Perkins, S. De Guise, S. Frasca, Jr. & R. A. French. 2004. Paramoebiasis associated with mass mortality of American lobster (Homarus americanus) in Long Island Sound, USA. J. Aquat. Anita. Health 16:29-38.

Mullen, T. E., K. R. Nevis, C. J. O'Kelly, R. J. Gast & S. Frasca, Jr. 2005. Nuclear small-subunit ribosomal RNA gene-based characterization, molecular phylogeny and PCR detection of the Neoparamoeba from western Long Island Sound lobster. J. Shellfish Res. 24(3):719-731.

Newman, M. W. & G. E. Ward, Jr. 1973. An epizootic of blue crabs, Callinectes sapidus, caused by Paramoeba perniciosa. J. Invert. Path. 22:329-334.

Page, F. C. 1973. Paramoeba: a common marine genus. Hydrobiologia 41:183-188.

Peglar, M. T., L. A. A. Zettler, O. R. Anderson, T. A. Nerad, P. M. Gillevet, T. E. Mullen, S. Frasca, Jr., J. D. Silberman, C. J. O'Kelly & M. L. Sogin. 2003. Two new small-subunit ribosomal RNA gene lineages within the subclass Gymnamoebia. J. Euk. Microbiol. 50:224-232.

Perkins, F. O. & M. Castagna. 1971. Ultrastructure of the Nebenkorper or "secondary nucleus" of the parasitic amoeba Paramoeba perniciosa (Amoebida, Paramoebidae). J. Invert. Pathol. 17:186-193.

Russell, S., K. Hobbie, T. Burrage, C. Koerting, S. De Guise, S. Frasca, Jr. & R. A. French. 2000. Identification of a protozoan parasite in the American lobster, Homarus americanus, from Long Island Sound. J. Shellfish Res. 19:581.

Saccone, C., C. Gissi, C. Lanave, A. Larizza, G. Pesole & A. Reyes. 2000. Evolution of the mitochondrial genetic system: an overview. Gene 261:153-159.

Saitou, N. & M. Nei. 1987. The Neighbor-joining Method: A New Method for Reconstructing Phylogenetic Trees. Mol. Biol. Evol. 4:406-425.

Scheibling, R. E. & A. W. Hennigar. 1997. Recurrent outbreaks of disease in sea urchins Strongylocentrotus droebachiensis in Nova Scotia: evidence for a link with large-scale meteorologic and oceanographic events. Mar. Ecol. Prog. Ser. 152:155-165.

Scott, J. 1995. A place in the world for RNA editing. Cell 81:833-836.

Sprague, V. & R. L. Becken. 1966. A disease of blue crabs (Callinectes sapidus) in Maryland and Virginia. J. Invert. Path. 8:287-289.

Sprague, V. & R. L. Beckett. 1968. The nature of the etiological agent of "gray crab" disease. J. Invert. Path. 11:503.

Sprague, V., R. L. Beckett & T. K. Sawyer. 1969. A new species of Paramoeba (Amoebida, Paramoebidae) parasitic in the crab Callinectes sapidus. J. Invert. Path. 14:167-174.

Stuart, K. D., A. K. Panigrahi & R. Salavati. 2001. RNA editing in kinetoplastid mitochondria. In: B. L. Bass, editor. Frontiers in molecular biology. Oxford: Oxford University Press. pp. 1-19.

Taylor, M. S. & M. E. Hellberg. 2003. Genetic evidence for local retention of pelagic larvae in a Caribbean Reef fish. Science 299:107-108.

Tedesco, M. & M. Van Patten. 2000. Lobster mortalities lead to fisheries disaster in Long Island Sound. Coastlines (EPA Publication) issue 10.4:1-13.

Utermohl, H. 1958. Zur Vervollkommnung der quantitativen Phytoplankton-Methodik. Mitt. int. Ver. Theor. Angew. Limnol 9:1-38.

Wong, F. Y. K., J. Carson & N. G. Elliot. 2004. 18S ribosomal DNA-based PCR identification of Neoparamoeba pemaquidensis, the agent of amoebic gill disease in sea-farmed salmonids. Dis. Aquat. Org. 60:65-76.

Zhang, H., D. Bhattacharya & S. Lin. 2005. Phylogeny of dinoflagellates based on mitochondrial cytochrome b and nuclear small subunit rDNA sequence comparisons. J. Phycol. 41:411-420.

Zhang, H. & S. Lin. 2002. Identification and quantification of Pfiesteria piscicida by using the mitochondrial cytochrome b gene. Appl. Environ. Microbiol. 68:989-994.

SENJIE LIN * AND HUAN ZHANG

Department of Marine Sciences, University of Connecticut, 1080 Shennecossett Road, Groton, Connecticut 06340

* Corresponding author: E-mail: senjie.lin@uconn.edu

COPYRIGHT 2005 National Shellfisheries Association, Inc.
COPYRIGHT 2005 Gale Group

Return to Mitochondrial Diseases
Home Contact Resources Exchange Links ebay