Michigan State University This is to certify that the dissertation entitled Molecular Epidemiology and Evolution of Sarcocystis neurona, The Etiologic Agent of Equine Protozoal Myeloencephalitis (EPM) presented by Hany M. Elsheikha has been accepted towards fulfillment of the requirements for the PhD. degree in Large Animal Clinical Sciences MSU is an Affirmative Action/Equal Opportunity Institution PLACE IN RETURN Box to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 c:/ClRC/DateDue.indd-p.15 —_— MOLECULAR EPIDEMIOLOGY AND EVOLUTION OF SARCOCYSTIS NEURONA, THE ETIOLOGIC AGENT OF EQUINE PROTOZOAL MYELOENCEPHALITIS (EPM) By Hany M. Elsheikha A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Large Animal Clinical Sciences 2004 ABSTRACT MOLECULAR EPIDEMIOLOGY AND EVOLUTION OF SARCOCYSTIS NEURONA, THE ETIOLOGIC AGENT OF EQUINE PROTOZOAL MYELOENCEPHALITIS (EPM) By Hany M. Elsheikha Equine Protozoal Myeloencephalitis (EPM) is a common, serious neurological disease of horses that can lead to disability and death. EPM is caused by the ubiquitous protozoan parasite Sarcocystis neurona, Horses acquire this parasite from the feces of opossums of the genus Didelphis when they ingest the sporocyst stage in contaminated food. Inadequate understanding of the parasite’s species identity, genetic makeup, host range, and the mechanisms it uses to cause disease in horses has hampered effective diagnosis and management of the disease. The most basic information needed is how to determine which Sarcocystis strains are S. neurona and which are not and how genetic traits are transferred within the population. The main hypothesis investigated was that S. neurona undergoes genetic recombination during the sexual phase of parasite development leading to strains with variable genetic traits that may include increased virulence. The studies described here included the assembly of a large collection of S. neurona and other related cyst-forming coccidia, development of methods for purification of the sporocyst stage from intestinal cells of opossums and the merozoite stage from tissue culture and the use of DNA sequencing of the small subunit ribosomal RNA (SSUrRNA) gene-and major surface antigen gene 1 (SAGI) as well as sequence of the 25/396 diagnostic marker. Amplified Fragment Length Polymorphism (AF LP) methodology was also used as a whole-genome fingerprinting method to develop a “species identity” or “genetic fingerprint” for S. neurona strains. This method was applied to characterize the available Sarcocystis strains that have been assessed by preliminary phenotypic and molecular marker methods such as PCR-RFLP as “neurona- like”. The contributions of the present work are: (1) evaluation of reproducible and sensitive methods for classification of strains of S. neurona from horses and opossums to the species and subspecies level based on SAGl gene sequence data and AF LP markers; (2) elucidation of the nature and extent of genetic variation and phenotypic heterogeneity of Sarcocystis spp. isolated from horses, opossums, and other vertebrates; and (3) use of phylogenetic information to assess the distribution of S. neurona in equids and wild hosts and to predict the risk of introduction into horses. Methods developed in this work can be used to identify to the species and subspecies level any S. neurona isolate, to track sources of S. neurona for horses, and to test whether S. neurona undergoes genetic recombination to produce new strains with pathogenic potential. DEDICATION With love, I dedicate my dissertation to My beloved wife, Mom, and the spirit of my Dad iv ACKNOWLEDGEMENTS Many people have made the completion of this work possible. First of all, I would like to thank my major advisor Dr. Linda S. Mansfield for the very productive ideas we shared. Dr. Mansfield offered me every possible way of support to accomplish this work. Dr. Mansfield was always a great guidance and constructive challenge for me. I also want to thank my co-major advisor Dr. Mahdi A. Saeed for always being an inspiring figure in my career and for his help to come to Michigan State University. I would also like to thank my committee members for their continuous and generous support and cooperation throughout this work. Thank you, Dr. Scott D. Fitzgerald. I thank you not only because you allowed me to conduct part of my dissertation research at the Animal Health Diagnostic Laboratory, but especially for the times in which you reached out to help as an advisor and as a friend. I also would like to thank Dr. Thomas S. Whittam for providing me with the privilege of carrying out part of my dissertation research in the Microbial Evolution Laboratory. His love of science and his dedication to his work inspire me. My sincere thanks to Dr. Harold C. Schott and Dr. Jeffrey P. Massey for their encouragement and enthusiasm and for providing me with guidance in my research and in my career. Many thanks to the past and present staff at the Parasitology Section, Diagnostic Center for Population and Animal Health; the Whittam and the Mansfield laboratory members; and most of all to Alice Murphy, David Lacher, Seth Walk, and Lindsey Ouellette, Adam Nelson, Geetha Parthasarathy, David Wilson, Dr. Julia Bell, Kathryn Jones, Lakeisha Cummingham, and Dr. Mary Rossano. I had great moments with all of them and very friendly relationships. I am also indebted to the Department of Large Animal Clinical Sciences staff for helping me in various ways during my stay in the department and for providing me with all the necessary advice and support to accomplish my academic goals. Financial support was offered by the Egyptian government; Department of Large Animal Clinical Sciences, College of Veterinary Medicine, Michigan State University; and the Grayson Jockey Club Research Foundation. I want to thank my wife Dr. Nashwa for her love and support in every possible way she could and for being next to me during this exciting time of my life. And of course my beloved mother, although a few thousand miles away, she has fully supported me and encouraged me in every step of my life. Finally, I wish to honor the memory of my father who I missed since my childhood but his spirit has always been next to me. vi TABLE OF CONTENTS LIST OF TABLES ........................................................................................................... ix LIST OF FIGURES ......................................................................................................... xi KEY TO ABBREVIATIONS ....................................................................................... xvi THE RESEARCH PROBLEM (SCOPE AND SIGNIFICANCE) ............................... 1 References ................................................................................................................ 4 HYPOTHESE STESTED .................................................................................................. 5 CHAPTER 1: LITERATURE REVIEW ........................................................................ 8 Life cycle of Sarcocystis neurona ............................................................................ 8 Sarcocystis neurona and Equine Protozoal Myeloencephalitis ............................... 9 Taxonomic status of S. neurona ............................................................................. 11 Previous genetic and phylogenetic studies of Sarcocystis neurona ....................... 12 Genetic variation among Sarcocystis neurona isolates ......................................... 16 References .............................................................................................................. 19 CHAPTER 2: PURIFICATION OF SARCOCYSTIS NE URONA MATERIALS ..... 24 [1] Purification of S. neurona sporocysts from opossum (Didelphis virginiana) using potassium bromide discontinuous density gradient centrifugation ............................... 25 Abstract .................................................................................................................. 26 Introduction ............................................................................................................ 27 Materials and Methods ........................................................................................... 29 Results .................................................................................................................... 36 Discussion .............................................................................................................. 39 References .............................................................................................................. 42 [2] Generally applicable methods to purify intracellular coccidia from cell cultures and to quantify purification efficacy using quantitative PCR .............................................. 44 Abstract .................................................................................................................. 45 Introduction ............................................................................................................ 46 Materials and Methods ........................................................................................... 48 Results .................................................................................................................... 57 Discussion .............................................................................................................. 71 References .............................................................................................................. 73 CHAPTER 3: PHYLOGENETIC RELATIONSHIPS OF SARCOCYSTIS NE URONA ISOLATES OF HORSE AND OPOSSUM TO OTHER CYST- FORMING COCCIDIA DEDUCED FROM SSUrRNA GENE SEQUENCES ........ 74 Abstract .................................................................................................................. 75 Introduction ............................................................................................................ 76 Materials and Methods ........................................................................................... 79 vii Results .................................................................................................................... 89 Discussion ............................................................................................................ 101 References ............................................................................................................ l 08 CHAPTER 4: SARCOCYSTIS NE URONA MAJOR SURFACE ANTIGEN GENE 1 (SA GI) SHOWS EVIDENCE OF HAVING EVOLVED UNDER POSITIVE SELECTION PRESSURE ............................................................................. - ............... 113 Abstract ................................................................................................................ 114 Introduction .......................................................................................................... l 15 Materials and Methods ......................................................................................... 117 Results .................................................................................................................. 121 Discussion ............................................................................................................ 130 References ............................................................................................................ 1 3 6 CHAPTER 5: PHYLOGENETIC CONGRUENCE OF SARCOCYSTIS NE URONA ISOLATES IN THE UNITED STATES BASED ON SEQUENCE ANALYSIS AND RESTRICTION FRAGMENT LENGTH POLYMORPHISM ................................ 138 Abstract ................................................................................................................ 139 Introduction .......................................................................................................... 140 Materials and Methods ......................................................................................... 142 Results .................................................................................................................. 149 Discussion ............................................................................................................ 163 References ............................................................................................................ 167 CHAPTER 6: GENETIC VARIATIONS AMONG SARCOCYSTIS NEURONA AS REVEALED BY AFLP MARKERS ............................................................................ 169 Abstract ................................................................................................................ 169 Introduction .......................................................................................................... 1 70 Materials and Methods ......................................................................................... 172 Results .................................................................................................................. l 78 Discussion ............................................................................................................ 183 References ............................................................................................................ 1 87 CHAPTER 7: DISCUSSION AND CONCLUDING REMARKS ............................ 189 DISCUSSION ...................................................................................................... 189 CONCLUSIONS .................................................................................................. 202 FUTURE DIRECTIONS ..................................................................................... 206 REFERENCES .................................................................................................... 212 viii LIST OF TABLES CHAPTER 2 [1] Purification of Sarcocystis neurona sporocysts from opossum (Didelphis virginiana) using potassium bromide discontinuous density gradient centrifugation Table 1. Comparison between mucosa] scraping (current method) and potassium bromide discontinuous gradient centrifugation method for the isolation of Sarcocystis spp. sporocysts from opossum small intestine ........................ 37 [2] Generally applicable methods to purify intracellular coccidia from cell cultures and to quantify purification efficacy using quantitative PCR Table 1. qPCR primers ............................................................................................. 55 Table 2. Total DNA concentration (ng/ul) in purified and unpurified preparations of DNA estimated to each contains 1000 merozoites/ZOOul estimated by UV absorbance spectroscopy ..................................................................... 62 Table 3. Total and parasite DNA content of defined admixtures of crude and column purified preparations of B. darlingi estimated by UV absorbance and qPCR targetting parasite rDNA .......................................................... 62 Table 4. Effect of column purification on the detection of host rDNA amplification in preparations estimated to contain 1000 merozoites/ZOOul .................... 70 Table 5. Effect of column purification on the detection of parasite rDNA amplification in preparations contain 1000 merozoites/ZOOul .................. 70 Table 6. Effect of column purification on the detection of parasite rDNA in lng/ul total DNA ................................................................................................... 70 CHAPTER 3 Table 1. Details of Apicomplexan taxa included in the study ................................. 80 Table 2. Names, locations, and DNA sequence of primers used for amplification and sequencing of portion of SSUrRNA gene from Sarcocystis neurona isolates from horses and opossums. ........................................................... 82 Table 3. The log likelihood scores (- 1n likelihood) of the topologies of ML, NJ, and MP trees of the SSUrRNA under best-fit models of nucleotide evolution selected by LRT and AIC testing using program modeltest. TrN = (Tamura and Nei 1993), GTR = (Yang, 1994). ...................................... 97 ix Table 4. Proportion of pairwise nucleotide differences between SSUrRNA sequences of Sarcocystis spp. and other organisms determined by Uncorrected ("p") distance analysis of alignment ...................................... 98 CHAPTER 4 Table 1. DNA sequence differences of SA G1 gene of S. neurona isolates based on pairwise comparisons using uncorrected ("p") distance matrix ............... 122 CHAPTER 5 Table 1. Details of the Sarcocystis species analyzed in the present study. ............ 143 CHAPTER 6 Table 1. Origin, source, year of isolation of taxa used in this study. ..................... 173 Table 2. Selected primer combinations and polymorphism rates for AF LP analysis of 15 Sarcocystis neurona strains. ........................................................... 179 CHAPTER 2 LIST OF FIGURES [1] Purification of Sarcocystis neurona sporocysts from opossum (Didelphis virginiana) using potassium bromide discontinuous density gradient centrifugation Figure 1. Figure 2. Figure 3. Centrifuge tubes showing a white band (arrows) of Sarcocystis neurona sporocysts afier potassium bromide discontinuous gradient centrifugation (A) at 3,000 g, 30 min, and 4°C. (B) At 16,000 g, 1 h, and 4°C ............... 32 Photomicrographs of cleaned and crude S. neurona sporocysts. Crude sporocysts were collected and cleaned from intestinal mucosa of opossum. (A, B) Phase contrast microscopy of crude and cleaned sporocysts. Scale bar = 25 ul (C) Oblique illumination of purified sporocysts. Note the residual body (white arrow), sporozoite (black arrow), oocyst wall (black arrowhead), and sporocyst wall (white arrowhead). Scale bar = 100 pl ...36 Results of random amplified polymorphic patterns using the primers JNBZS/JD396 and restriction enzyme digestion with Hinfl and HindIII of sporocysts DNA preparations from Sarcocystis neurona isolates from Michigan opossums (MIOP), S. neurona (MIHl) isolate from horse, and S. falcatula (Cornell) isolate from opossum. M IOO-bp ladder marker; P PCR product; Hf Hinfl digest; Hd HindIII digest. Note, after HindIII digestion of PCR 334 bp amplicons, S. neurona amplicon is unaffected while S. falcatula amplicon yields bands at 164, 108, and 62 bp. After Hinfl digestion, S. neurona amplicon yields bands at 180 and 154 bp while S. falcatula amplicon yields bands at 170 and 164 bp. .............................. 38 [2] Generally applicable methods to purify intracellular coccidia from cell cultures and to quantify purification efficacy using quantitative PCR Figure 1. Figure 2. Figure 3. Illustration of the filter assembly. It consisted of a removable 250-ml borosilicate glass funnel (l) and support base (2). An anodized aluminum spring clamp (3) sandwiched a cellulose filter pad or filter paper (4) between funnel and support base. Connection to SOO-ml Erlenmeyer filter flask (5) is made with a silicone stopper (6). Parasite suspension with debris (7) and relatively pure parasite suspension (8) ............................... 50 Photomicrographs of S. neurona merozoites. (A) Crude merozoite preparations before purification. Scale bar = 50 ul. (B) Merozoites (m) completely free of host cell (hc) contaminants and extraneous debris after purification using PD-lO column. Scale bar = 25 ul ................................. 57 Purification of Sarcocystis neurona merozoites (A) and Besnoitia darlingi tachyzoites (B) cultured in Bovine turbinate cells using PD-10 columns, filter pads, and filter papers compared to unpure (crude) merozoites. xi Number of parasites is expressed as x 106. All experiments were conducted in triplicate ................................................................................ 59 Figure 4. Silver-stained SDS-PAGE gel of whole-cell protein preparations of Sarcocystis neurona merozoites showing differences in the protein profiles of crude S. neurona merozoites cultured in BT cells and PD-10 column- purified merozoites. M, molecular weight markers (Broad Range Standards; Bio-Rad); Lanes 1 to 3 contain the following amounts of S. neurona protein extract of column- purified merozoites: 1, 2p]; 2, Sul; 3, 10ul. Lane 4 contains 15p] of S. neurona protein extract of un-pure merozoites. The numbers on the left indicate the molecular masses of proteins in kilodaltons (KDa) used as standard markers ........................... 60 Figure 5. Analytical sensitivity of PCR analyses using the diagnostic primers JNB25/J D396 on DNA Obtained from crude Sarcocystis neurona merozoites cultured in BT cells and PD-lO column-purified S. neurona merozoites. Lanes: M, molecular size marker of a 100bp ladder; 1, 2, 3 unpure merozoites (1000, 100, 10, respectively); 4, 5, 6 column- purified merozoites (1000, 100, 10, respectively); 7, a positive control; 8 a negative control. Numbers on the right and on the left of the gel are DNA fragment sizes (in base pairs). ................................................................................... 63 Figure 6. Amplification plot (fluorescence vs. cycle number) in quantitative PCR assays. (A) Proportional delay in the amplification of parasite rDNA as crude B. darlingi are diluted, in serial 5-fold steps, over 4 orders of magnitude. (B) After normalizing crude (red) and column-purified (blue) extracts to a total DNA concentration of lng/ul, amplification of host 18S rDNA is delayed. (C) Note amplification of B. darlingi l8S rDNA is accelerated. (D) Note amplification of B. darlingi hsp70 is accelerated ...66 CHAPTER 3 Figure 1. TEM of S. neurona merozoites cultured in equine dermal (ED) cells and budding out of the residual body (RB). They had conoids (CO), micronemes (MN), and dense granules (DG). Note the presence of dividing nucleus (N) and nucleolous (NO) and the absence of rhoptries, which exist in S. falcatula. Scale bar = 1pm ............................................. 89 Figure 2. Strict consensus of 117 most parsimonious trees inferred from analysis of 347 phylogenetically informative characters of the nuclear SSUrRNA gene of 17 Sarcocystis spp. and 11 related taxa and rooted with three Outgroup taxa. Neighbor-joining (NJ) tree topology based on Tajami Nei model was almost consistent with the MP trees. Values above and below the nodes represent bootstrap resampling results (% of 100 replications for MP analysis) and (% of 1000 replications for NJ algorithm), respectively. Numbers at the nodes ((1) indicate decay indices. Bootstrap values are reported only for clades present in >50% of replicates. GenBank xii Figure 3. Figure 4. CHAPTER 4 Figure 1. Figure 2. Figure 3. accession numbers of the SSUrRNA sequences of the organisms are given in parentheses. To the right taxonomic affiliation is presented. Branch lengths are proportional to distance. 50% majority role tree has identical topology ..................................................................................................... 94 Maximum likelihood phylogeny of SSUrRN A among the Sarcocystidae, rooted with three outgroup taxa constructed with a GTR + I + G model of substitution, (-1nL: 7885) heuristic search, random stepwise addition, TBR, equal character weight, with MULPARS in effect. The log likelihood score of the phylogeny is —7884.844. Numbers above nodes represent percent bootstrap support of 100 non-parametric replications. Branch lengths are proportional to the hypothesized amount of inferred evolutionary change (ML distances), as shown by the scale bar. Sequences of the isolates in bold are obtained in this study. Sequences of the isolates in bold are obtained in this study ............................................................... 96 Phylogram based on SSUrRNA sequence alignment of 31 Sarcocystidae taxa. Tree was constructed using the Bayesian inference and rooted on E. tenella, C. cayetanensis, and C. bigenetica. Posterior probabilities are displayed at each node. Sequences of the isolates in bold are obtained in this study .................................................................................................. 100 Maximum parsimonious phylogenetic tree from analysis of 208 phylogenetically informative characters of SA GI showing relationships among the Sarcocystis neurona isolates (rooted to T oxoplasma gondii). The first of the two numbers at the nodes represent bootstrap resampling results based on maximum parsimony analysis (MP, % of 1,000 replicates). The second number represents the bootstrap support using neighbor-joining algorithm (NJ, % of 1,000 replicates). Numbers in parentheses after taxon names refer to GenBank accession nos., followed by the host and origin of isolate. Tree statistics are length (L)=290, consistency index (CI)=O.997, retention index (RI)=0.998 excluding uninforrnative characters .......................................................................... 123 Diversity in SA GI gene of Sarcocystis neurona. Plot of the variable nucleotide sites (A) and amino acid sites (B) in pairwise comparisons of the two alleles of SA GI gene with percentage of the nucleotide polymorphism shown above the domains. General three-domain structure model of Sarcocystis neurona SA GI gene is drawn above with the intron region shaded in grey. The arrow marks the signal peptide sequence region. Each breakpoint in the domains denotes the location of an alignment gap. Each vertical line marks the location of a nucleotide or amino acid difference between the two alleles sequence ......................... 126 Nucleotide alignment of 2 SA GI sequences representing the 2 alleles of S. neurona SA G] that shows predicted cleavage site for a signal peptide xiii (underlined) as identified by Hyun et al., (2003). Dots indicate sequence identity and dashes indicate alignment gaps ............................................ 127 Figure 4. Plot of the number of substitutions per 100 sites for synonymous (pg) and nonsynonymous (pN) sites between H1 and H2 SA G1 alleles in a 30-codon sliding window. The difference (pN_ps) is a measure of the level of selective constraint on various parts of the molecule .............................. 128 Figure 5. Selective pressures acting upon single codon sites of the SA GI gene using Approximate Likelihood Ratio (ARS) analysis. ...................................... 129 CHAPTER 5 Figure 1. Map of the United States illustrating the number and distribution of S. neurona isolates analyzed from different animal hosts in different states in the present study. ..................................................................................... 150 Figure 2. Nucleotide sequence alignment of the 25/396 marker sequences of Sarcocystis neurona isolates compared with other Sarcocystis spp. and - coccidian taxa. Sequences were truncated to the length of published sequences obtained from opossums from Florida and Mississippi. Bases that are identical to those of the prototype sequence are indicated by periods, missing bases are indicated by hyphen, and bases that are different from those of the prototype sequence are shown. Note the presence of 2 inserted bases (TA at positions 270 and 271) in Neospora caninum and in 10 out of 11 of Michigan S. neurona isolates ................ 153 Figure 3. Neighbor-joining (NJ) phylogenetic tree of Sarcocystis neurona isolates from various hosts and different localities in the US inferred from sequences of the 25/239 marker using General Time Reversible (GTR) model. One Sarcocystis sp. isolate (GenBank Accession no. AF 323943) was used as outgroup. Bootstrap confidence values were based on 1,000 replicates; only those values that exceed 60% are shown above the node. Branches without bootstrap values represent polytomies that could not be resolved because the sequences were almost identical. The scale bar represents the calculated distance value. Branch lengths are drawn proportional to the amount of genetic change. The isolate designation and locality follows taxon. For details of isolate designation please refer to Table 1 ..................................................................................................... 157 Figure 4. Maximum parsimony (MP) phylogenetic tree of Sarcocystis neurona isolates from various hosts and different localities in the US inferred from sequences of the 25/239 marker ............................................................... 158 Figure 5. Neighbor-joining (NJ) phylogenetic tree of Sarcocystis neurona isolates from various hosts and different localities in the US inferred from xiv Figure 6. Figure 7. CHAPTER 6 Figure 1. sequences of the 25/239 marker using Kimura two-parameter distance model ........................................................................................................ 159 Phylogenetic tree of Sarcocystis neurona isolates obtained by minimum evolution (ME) method ............................................................................ 160 Diversity revealed by RF LP analyses of PCR products of representative isolates of S. neurona from Michigan. (A) RF LP pattern obtained by restriction digestion of the 25/396 marker (334 bp amplicons). Lanes 1 to 8, 334 bp undigested amplicons; lanes 9 to 16, samples treated with Hinfl; lanes 17 to 24, samples treated with HindIII. (B) RF LP pattern obtained by restriction digestion of the 33/54 marker (1100 bp amplicons). Lanes 1 to 8, 1100 bp undigested amplicons; lanes 9 to 16, Hian restriction profile; lanes 17 to 24, DraI restriction profile. Lanes M, molecular size standards are shown to the leftmost and rightmost of the figure in 100 bp-sized bands; (V) denote empty lanes. Sources of isolates of the amplicons and restriction analyses: MIH7 (lanes 1, 9, and 17), MIHlO (lanes 2, 10, and 18), MIOPl (lanes 3, 11, and 19), MIOP7 (lanes 4, 12, and 20), MSUOP/CBS (lanes 5, 13, and 21), MSUOP/CB6 (lanes 6, 14, and 22), MSUOP/cat2 (lanes 7, 15, and 23), MSUOP/Raccoonl (lanes 8, l6, and 24) ............................................................................................................ 162 An UPGMA dendrogram based on AF LP markers obtained with 9 primer pairs. Numbers shown at different nodes represent percentage confidence limits obtained in the bootstrap analysis. Nodes without numbers had bootstrap values of less than 50. The scale shown above is the measure of genetic similarity calculated according to J accard similarity coefficient prepared from AF LP banding patterns of 15 S. neurona strains. Taxon names and designations are indicated on the right side of the panel. ...... 181 XV AFLP ANOVA ATCC BT CSF Ct CV-l DMEM ED EMEM EPM HBSS Kbr ML MP NJ PBS PCR RAPD SAGl SSUrRNA UPGMA KEY TO ABBREVIATIONS Amplified fragment length polymorphism Analysis of variance American Type Culture Collection Bovine turbinate cell line Cerebrospinal fluid Cycle at which fluorescence significantly exceeds baseline levels African green monkey (Cercopithecus aethiops) kidney cells Dulbecco’s minimal essential medium Equine dermal cell line Eagle’s minimal essential medium Equine protozoal myeloencephalitis Hanks’ balanced salt solution Potassium bromide Maximum Likelihood Maximum Parsimony Neighbor Joining Phosphate buffered saline Polymerase chain reaction Random amplified polymorphic DNA Maj or surface antigen genel Small subunit ribosomal RNA gene Unweighted Pair Group Method with Arithmatic Mean xvi THE RESEARCH PROBLEM (SCOPE AND SIGNIFICANCE)- Equine protozoal myeloencephalitis (EPM) is a serious neurological disease of horses in the Americas caused by the protozoal parasite, Sarcocystis neurona (Dubey et al., 2001b). Over 50% of the horses in the US have been infected with S. neurona based on finding antibodies against the parasite in their blood (Bentz et al., 1997; Blythe et al., 1997; MacKay et al., 2000; Dubey et al., 2001b; Bentz et al., 2003; Rossano et al., 2003). The potential risk factors and characteristics associated with transmission of this pathogen are not fully understood. Several very closely related Sarcocystis organisms have been identified as S. neurona based on molecular procedures. However, these organisms have been shown to have different biological characteristics. Even though they all infect the opossum as the definitive host (with sexual stages), they do not all infect the same intermediate hosts (with the asexual stages). Also, they do not all produce disease in the gamma interferon gene knockout mice (y-IFN-KO) that have been used to mimic infection in the horse (Cheadle et al., 2001; Dubey etal., 2001a, 2001b). These findings suggest that they vary in important ways that are likely to affect the type of disease produced or even whether they are able to produce disease in horses. Veterinarians have observed that not all horses exposed to the parasite develop the disease, and when disease exists, clinical signs vary from mild ataxia to paralysis, recumbency, and death. Also, different horses have different responses to treatment, suggesting that strain variations in the parasite may be an important factor in the treatment of EPM. Strain variation may contribute to problems in diagnosis of EPM and will certainly affect any vaccine that is made to protect horses. With the increasing awareness for the potential of S. neurona as an important cause of economic losses in the horse industry, knowledge of the mechanisms by which it causes disease are of great importance. In this regard, understanding the extent and nature of genetic variation in S. neurona and closely related species is an essential prerequisite to determining the epidemiology and transmission of this pathogen. It is especially important to know whether genotypes of S. neurona are stable (indicative of clonal propagation) or unstable (due to frequent genetic recombination). For example, in a related protozoan T oxoplasma gondii, population studies demonstrated that this organism is genetically stable, and the small amount of variation defined three clonal lineages where phenotypes such as virulence are associated with a single lineage (Howe and Sibley, 1995). This situation may be due to the fact that only one isolate of T oxoplasma is usually found in a cat at a given time; therefore, the potential for sexual recombination (which can only occur in the host in which sexual reproduction takes place) is low. Other protozoa have high potential for sexual recombination, which generates diversity and enhances virulence attributes (Tibayrenc and Ayala, 2002; Mallon et al., 2003). The population structure of S. neurona has not been determined yet. However, the molecular tools and knowledge supporting the possibility of sexual recombination in this parasite are now available. It is well known that sexual reproduction of S. neurona takes place in the gut of opossums. Also, the recovery of many Sarcocystis strains from the gut of a single opossum shows that they co-exist and could undergo recombination (Elsheikha et al., 2004). Therefore, to test adequately the “recombination model” or “non-clonal model” theory for S. neurona and to determine population structure, a population genetic study was conducted in which data were analyzed in terms of allelic and genotypic frequencies, rather than simply assessing genetic distances as was done in the past. The objectives of present research were (1) to determine the population structure, phylogenetic relationships, and extent of genetic diversity among Sarcocystis spp. ' infecting horses and opossums in relation to closely related species using sequencing of multiple genetic markers and the Amplified Fragment Length Polymorphism (AF LP) methods that provide a genetic fingerprint and (2) to use this knowledge for thorough understanding of the molecular epidemiology and evolution of S. neurona at the molecular level. The specific aims of this work were: (1) to develop an accurate, reproducible, reliable and sensitive method for genotyping of clinical strains of S. neurona from horses and environmental strains from opossums to subspecies level based on molecular markers; (2) to elucidate the nature and extent of genetic variation as well as the heterogeneity of Sarcocystis species isolated from horses and opossums; (3) to improve the systematic classification of S. neurona strains according to their evolutionary relationships; (4) to use phylogenetic information to predict the pattern of distribution of S. neurona among equines and wild host species; and (5) to refine the definition of S. neurona species, strains, clones, and populations using molecular evolutionary analysis. This work was focused on S. neurona, but the genetic fingerprints of S. neurona isolates were compared to diverse but related organisms such as S. falcatula, B. darlingz’, T. gondii, and N. caninum. The work appears to be the first wide—ranging comparative study of S. neurona and related species from domesticated and wild animal hosts. REFERENCES Bentz BG, Granstrom DE, Stamper S (1997) Seroprevalence of antibodies to Sarcocystis neurona in horses residing in a county of southeastern Pennsylvania. J Am Vet Med Assoc 210:517—518 Bentz BG, Ealey KA, Morrow J, Claypool PL, Saliki JT (2003) Seroprevalence of antibodies to Sarcocystis neurona in equids residing in Oklahoma. J Vet Diagn Invest 62597—600 Blythe LL, Granstrom DE, Hansen DE, Walker LL, Bartlett J, Stamper S (1997) Seroprevalence of antibodies to Sarcocystis neurona in horses residing in Oregon. I Am Vet Med Assoc 210:525-527 Cheadle MA, Tanhauser SM, Scase TJ, Dame JB, Mackay RJ, Ginn PE, Greiner EC (2001) Viability of Sarcocystis neurona sporocysts and dose titration in gamma- interferon knockout mice. Vet Parasitol 95:223—231 Dubey JP, Lindsay DS, Kwok OC, Shen SK (2001a) The gamma interferon knockout mouse model for Sarcocystis neurona: comparison of infectivity of sporocysts and merozoites and routes of inoculation. J Parasitol 872117—1173 Dubey JP, Lindsay DS, Saville WJA, Reed SM, Granstrom DE, Speer CA (2001b) A review of Sarcocystis neurona and equine protozoal myeloencephalitis (EPM). Vet Parasitol 95289—1 31 Elsheikha HM, Murphy AJ, Mansfield LS (2004) Prevalence of Sarcocystis species sporocysts in Northern opossums (Didelphis virginiana). Parasitol Res 93:427—431 Howe DK, Sibley LB (1995) T oxoplasma gondii comprises three clonal lineages: correlation of parasite genotype with human disease. J Infect Dis 172:561—1566 MacKay RJ, Granstrom DE, Saville WJ, Reed SM (2000) Equine protozoal myeloencephalitis. Vet Clin North Am Equine Pract 16:405—425 Mallon M, MacLeod A, Wastling J, Smith H, Reilly B, Tait A (2003) Population structures and the role of genetic exchange in the zoonotic pathogen Cryptosporidium parvum. J Mol Evol 561407—417 Rossano MG, Kaneene JB, Marteniuk JV, Banks BD, Schott HC 2nd, Mansfield LS (2003) A herd-level analysis of risk factors for antibodies to Sarcocystis neurona in Michigan equids. Prev Vet Med 57:7—13 Tibayrenc M, Ayala FJ (2002) The clonal theory of parasitic protozoa: 12 years on. Trends Parasitol 18:405—41 0 HYPOTHESES TESTED Whether in nature the parasitic protozoan Sarcocystis neurona reproduces sexually or clonally remains largely unsettled. This ignorance may seem surprising given that the matter is of great medical consequence: the strategies for developing vaccine and curative drugs as well as for diagnosis and treatment are different for an organism that propagates clonal, than for an organism that propagates with frequent sexual recombination. Overall hypothesis The overall hypothesis tested in this work was that: (H0) S. neurona has panamictic genetic structure due to frequent genetic recombination during the sexual phase of parasite development in the intestinal cells of the definitive host leading to strains with variable genetic traits including increased virulence. An alternate hypothesis (H1) was that S. neurona has a clonal population structure and virulence is associated with a single lineage. Testing of this hypothesis can provide a clearer understanding of the population structure and transmission characteristics of this ubiquitous parasite. Specific hypotheses CHAPTER 3 Hypothesis 1: S. neurona strains from opossums and horses form a monophyletic group. Ho: S. neurona strains form a single monophyletic group. H]: S. neurona strains are not monophyletic. Hypothesis 2: Phylogenies of S. neurona and its hosts are consistent with a common history. Ho: Associated hosts and parasites have identical phylogenies. H1: There are different phylogenies for hosts and parasites. Rationale: These two hypotheses were investigated by determining the phylogenetic relationships of S. neurona strains obtained from horses and opossums in relation to other cyst-forming coccidia based on sequences of the small subunit ribosomal RNA (SSUrRNA) gene. CHAPTER 4 Hypothesis 3: SAGl gene of S. neurona evolved under positive selection pressure. Ho: The SAGl gene of S. neurona has evolved under diversifying selection pressure. H1: The SAGl gene has not evolved under positive selection pressure. Rationale: Sequences of the SAGl gene of multiple S. neurona strains were used to investigate this hypothesis. CHAPTER 5 Hypothesis 4: S. neurona strains have a clonal population structure. Ho: S. neurona strains exhibit phylogenetic congruence and clonality. H1: S. neurona strains include diverse, distinct genotypes. Rationale: Sequences of the diagnostic 25/396 marker were obtained from multiple S. neurona strains from different hosts in the United States and were used to explore this hypothesis. CHAPTER 6 Hypothesis 5: A genome-wide fingerprinting analysis using AF LP resolves the intra- species relationships among S. neurona. Ho: AF LP resolves the intraspecific relationships among S. neurona. H1: AF LP does not resolve the intraspecific relationships among S. neurona. Rationale: To investigate this hypothesis, the phylogenetic relationships between S. neurona strains were constructed based on fingerprinting profiles of S. neurona strains obtained by the AF LP technique. CHAPTER 1 LITERATURE REVIEW Life cycle of Sarcocystis neurona. Sarcocystis species have an obligatory, two- host life cycle (F ayer et al., 1982; Dubey et al., 1989a). Sarcocystis species replicate asexually in the muscles or other tissues of a prey intermediate host, where they form tissue cysts that are infective to an appropriate carnivorous definitive host. The parasite reproduces sexually in the intestine of the definitive host, ultimately producing sporocysts that are passed in feces. Ingestion of these sporocysts by an intermediate host species continues the life cycle. Opossums have been identified as the natural definitive host of Sarcocystis neurona (Fenger et al., 1995, 1997). A recent study revealed that domestic cats can serve as intermediate hosts for S. neurona when they are experimentally infected with S. neurona sporocysts obtained fi'om naturally infected opossums (Dubey et al., 2000). Tissue cysts produced in these cats were infective to opossums and were subsequently confirmed to be S. neurona. Also, numerous S. neurona strains have been isolated from various geographic regions and biological hosts in the US, including nine- banded armadillos, skunks, raccoons, sea otters, harbor seals, and cats (Dubey et al., 2000; Dubey et al., 2001b; Cheadle et al., 2001b, 2001c; Tanhauser et al., 1999). Horses are infected with the parasite when they ingest sporocysts in food contaminated with opossum feces (Fenger et al., 1995). The sequence of events that occurs when the parasite enters the horse is not known. It has been speculated that in the horse the parasite enters the blood vessels from the gut and then invades the central nervous system (rather than muscle), destroying cells as it multiplies. The only stage of the parasite found to date in horses is the asexual merozoite stage in neural cells. S. neurona sarcocysts are absent fi'om other horse tissues suggesting that horses are an accidental or “dead-end” host for the parasite. As such, an infected animal cannot transmit the parasite to another horse or back to the opossum. The absence of S. neurona sarcocysts in horses raised some concerns regarding the authenticity of horses as a true intermediate host (I.H). Sarcocystis neurona and equine protozoal myeloencephalitis (EPM). S. neurona is the most commonly diagnosed neurological disease EPM in the horse population in North America (Dubey et al., 2001 a). Although other coccidia have been also associated with the distinct clinical manifestations of EPM, only S. neurona is well known to be responsible for causing EPM disease. Approximately 50% (20 to 90%) of horses in the US have been infected by this organism as evidenced by antibodies against the parasite in their serum (Bentz et al., 1997; Blythe et al., 1997; MacKay et al., 1997, 2000; Bentz et al., 2003; Rossano et al., 2003). Market research conducted with equine veterinarians revealed that EPM was diagnosed in 220,000 horses in 1998 with a conservative estimate of 10,000 to 160,000 new cases every year. Approximately 70% of diseased horses improved with therapy, but less than one half returned to normal function. EPM is a serious, growing problem for the horse industry in the US. Animals affected by EPM can demonstrate a variety of clinical abnormalities, and signs can vary in severity from mild lameness to death. EPM-affected horses are unable to attain their ultimate performance. Despite treatment, horses with EPM can be affected so severely that they must be euthanatized, or they can develop permanent disabilities, which can affect their athletic activities. Treatment of EPM is expensive, and both mildly and severely affected horses often require treatment for extended periods without guarantee of success (MacKay et al., 2000; Dubey et al., 2001a). Economists from the USDA estimated that the average cost to the equine industry in the US due to EPM, is $27 million annually (Dubey et al., 2001a). This figure includes estimates for costs related to diagnosis, treatment, and maintenance of horses while they are unable to be used for their intended purpose, loss of stakes payments, and the value and transport costs of horses that die or are euthanatized. This estimate does not account for opportunity losses because affected horses are unable to attain their ultimate performance potential. The National Animal Health Monitoring System’s Equine ‘98 study rated “Gathering information about this disease (EPM)” as one of the highest priorities for the equine industry (NAHMS, 1998). Veterinarians throughout the US have diagnosed this disease with increasing frequency since commercial tests to detect S. neurona infection have become available. However, conventional diagnostic methods for EPM have suffered from problems with false positive tests. That is, normal non- neurological horses often have S. neurona antibodies in serum and cerebrospinal fluid. Sarcocystis neurona was first isolated and characterized in 1991 by Dubey and colleagues. Even today there are few definitive morphological characteristics with which to identify Sarcocystis taxa that infect horses. Also, there have been difficulties in characterizing S. neurona strains cultured in the laboratory in order to clearly identify species. At least four species of Sarcocystis have been described in the opossum including S. neurona (Dubey et al., 1991), S. falcatula (Box and Duszynski, 1980), S. speeri (Dubey and Lindsay, 1999), S. lindsayi (Dubey et al., 2001c), and each species has been considered in turn as a possible causative agent of EPM in horses. Recently, another Sarcocystis sp. that cycles between opossum and the brown-headed cowbird (Molothrus ater) was identified and found to causes neurological disease in the mouse 10 model of EPM (Mansfield et al., unpublished data). All of these species have sexual stages (sporocysts) in the opossum gut that cannot be easily distinguished morphologically in fecal specimens or at necropsy, and this further complicates the determination of parasite diversity. Variation between isolates has been observed for some biological and molecular characters. Not all horses exposed to the parasite develop the disease; but when disease exists, clinical signs can vary from mild ataxia to paralysis, recumbency, and death (MacKay et al., 2000). Parasite diversity may also be reflected in variable responses to treatment in horses infected with S. neurona suggesting that strain-to-strain variation may be a factor in the clinical management of EPM. The fact that horses are exposed to multiple Sarcocystis spp. in the environment may complicate the understanding of the life cycle of S. neurona and of the pathogenesis and diagnosis of EPM. Clearly, additional features must be found to aid in the identification and characterization of Sarcocystis spp. The application of sensitive molecular approaches has obviated the need for expensive, cumbersome, and time-consuming biological infection studies as the sole means of distinguishing species. Taxonomic status of Sarcocystis neurona. The authors of early studies named and described different Sarcocystis spp. on the basis of host occurrence and the appearance of the muscle sarcocysts (Wenvon, 1926; Boch et al., 1979). Following the discovery in 1972 of the obligate heteroxenous (requiring two hosts) life cycle for the parasite, species were described on the basis of both intermediate and definitive host specificity (Heydron and Matuschka, 1981). Then, reports started to identify and to classify S. neurona mainly on phenotypic characteristics such as ultrastructural analysis, 11 host specificity (the ability to replicate/cause disease in certain hosts), details of the life cycle, or location of the developmental stages. Accurate identification of a Sarcocystis sp. using solely morphological criteria may not be adequate, yet distinguishing the true host for a given Sarcocystis sp. from related Sarcocystis is critical to understanding and controlling of infection. Indeed, a new system of nomenclature was proposed for those species in which both host species were known (Heydron et al., 1975). However, cross- transmission studies have indicated that the host specificity of certain species is not as strict as previously thought. Furthermore, although certain morphological and ultrastructural characteristics of mature muscle cysts have been used to identify species (Mehlhom, 1976), some studies have revealed slight morphological heterogeneity occurring among cysts presumed to have belonged to single species. The identification of S. neurona has relied on biological infection patterns in a variety of animals. Biotyping based on susceptibility of interferon gamma knockout mice to S. neurona infection and the pattern of grth in tissue culture were the first methods used to differentiate between S. neurona, S. falcatula, and S. speeri. Whereas S. neurona and S. speeri sporocysts infect immunodeficient mice but not budgigars, the reverse is true for parasites designated S. falcatula (Marsh et al., 1997a; 1997b; Dubey and Lindsay, 1998). However, bioassay and development patterns in tissue culture systems are time- consuming and thus not-very suitable tools for systematic and epidemiological studies. Previous genetic and phylogenetic studies of S. neurona. Phylogenetic analysis can provide a quantitative estimate of genetic relatedness among S. neurona strains and patterns of divergence between members of this species and other Sarcocystis species. In the past decade, various methods have been developed for the identification and typing of 12 S. neurona strains. The first genetic analysis of S. neurona was done using a random amplified polymorphic DNA assay to compare S. neurona to Sarcocystis spp., T oxoplasma gondii, and several Eimeria spp. (Granstrom et al., 1994). Subsequently, the nuclear small subunit-ribosomal SSUrRNA gene was amplified from cultured S. neurona organisms, cloned, and sequenced (Fenger et al., 1994). Sequence information was used to design a PCR assay to screen DNA extracts from sporocysts shed by various candidate definitive hosts. The SSUrRNA gene from sporocyst DNA was cloned and sequenced and found to be identical to the SSUrRNA gene sequence of culture-derived S. neurona. Following the identification of the opossum as the definitive host (F enger et al., 1995), the SSUrRNA gene sequence of S. falcatula, which cycles between opossums and various birds, was determined (Fenger et al., 1994). While the sequences were identical, there is increasing evidence that suggests that Sarcocystis isolates from opossums were not a single uniform species, and subsequent biologic, morphologic, and molecular evidence showed that S. neurona and S. falcatula were distinct species (Dame et al., 1995; Marsh et al., 1999; Cheadle et al., 2001a). Phylogenetic analyses of the SSUrRNA gene from members of the family Sarcocystidae have been used to study relatedness and to help complete parasite life cycles. Initially, it was believed that protozoa co-evolved with the definitive host (Barta, 1989). The SSUrRNA gene sequence of S. neurona and results from experimental infection studies placed it in the felid clade of Sarcocystis (F enger et al., 1994). A recent analysis using SSUrRNA gene sequences from many more Sarcocystis spp. suggested intermediate host (5) were involved. S. neurona was placed in the non-ruminant clade and was found to be most closely related to S. mucosa, a parasite believed to use the 13 Bennetts wallaby (Dayrus sp.) as a definitive host (Jakes, 1998; Jenkins et al., 1999). Recently, molecular data in the form of DNA sequence of the small subunit ribosomal RNA genes (SSUrRNA) and the D2 domain of the large subunit ribosomal RNA genes have been used to reconstruct phylogenies and to examine the relationships of S. neurona and other Sarcocystis spp. (Slapeta et al., 2003). In spite of a wide array of analytical methods being available, the characterization of S. neurona strains below the species level remains a demanding task. Generally the lack of discriminatory power of serological methods greatly limits their applicability to the typing of S. neurona. Protein profiling methods such as gel electrophoresis and immunoblotting (Granstrom et al., 1993) also possess little discriminatory power and are time-consuming, require large samples for analysis, and may be difficult to interpret. Various DNA-based methods have been used for interspecies differentiation of Sarcocystis strains from horse and opossum. They include PCR-mediated restriction fragment length polymorphism and arbitrarily primed PCR (Tanhauser et al., 1999). However, these methods differ in their taxonomic range, discriminatory power, reproducibility, and ease of interpretation. In addition, these genetic typing assays also have drawbacks in that they require a relatively large amount of high-quality DNA and are difficult to standardize between laboratories. Thus, the need for high-resolution analytical tools, which will facilitate the typing of S. neurona on a routine basis, still exists. The ideal genotyping method produces results that are invariable from laboratory to laboratory and allows unambiguous comparative analyses and the establishment of reliable databases. One of the newest and most promising methods is amplified fragment length polymorphism (AF LP) analysis (Blears et al., 1999; V05 et al., 1995), developed 14 by Keygene BV, Wageningen, The Netherlands. This method combines universal applicability with high powers of discrimination and reproducibility. An increasing number of reports describe the use of AF LP analysis for animal genetic mapping, medical diagnostics, phylogenetic studies, and microbial typing (Blears et al., 1999). For example, AF LP analysis revealed a remarkably high degree of genetic diversity within Cryptosporidium parvum strains (Blears et al., 2000). AF LP is a whole-genome fingerprinting method based on selective amplification of restriction fragments (Vos et al., 1995; Blears et al., 1999). The AF LP reaction is a multistep procedure, which combines the power of PCR with the inforrnativeness of restriction enzyme analysis. The procedure includes the preparation of an AF LP template: genomic DNA digested with two restriction endonucleases that produce cohesive fragment ends and that cut DNA with different frequencies (rare cutter [RC] and frequent cutter [FC]). Following digestion, genomic restriction fragments are modified by ligation of synthetic adapters (RC adapter and F C adapter) with ends complementary to those of the restriction fragments. Thus, after the ligation step, genomic restriction fragments have termini with known sequences. Such an AF LP template is submitted to a highly stringent PCR amplification with primers fully complementary to their targets (adaptor and restriction site sequence), with additional nucleotides at the 3’ end, are used as selective oligomers to amplify a subset of ligated fragments. Only those fragments with complementary nucleotides extending beyond the restriction site are amplified by the selective primer under stringent annealing conditions. Fluorescent dye-labeling one of the AF LP primers (usually the RC primer) allows the detection of only a subset of the hundreds of amplified restriction fragments. Polymorphisms are identified by the 15 presence or absence of DNA fragments after separation on polyacrylarnide gels or a DNA sequencer. The technique has several advantages over other DNA fingerprinting systems. AF LP usually yields more complex banding patterns than most of the available DNA fingerprinting methods, which may increase discrimination between strains under analysis. The strategy of using two restriction enzymes and selective amplification provides a great flexibility in designing the typing protocols optimal for the given species. The most important of these are the capacity to inspect an entire genome for polymorphism without prior knowledge of the target DNA sequence and the reproducibility of the method. These features, combined with the automation and a high- throughput analysis, make AF LP superior to the currently used molecular techniques. In the present study, an AF LP-based procedure suitable for the inter- and intraspecies differentiation of S. neurona species of clinical and environmental origin was evaluated. Genetic variation among Sarcocystis neurona isolates. There is little information regarding genetic variation of S. neurona isolated from horses with EPM and other Sarcocystis spp. isolated fi‘om opossums. Recent studies of S. neurona support the evidence of isolate diversity obtained by phenotypic and western blot analysis of merozoite proteins (Mansfield et al., 2001; Marsh et al., 2001). S. neurona is differentially distributed in multiple mammalian hosts and may be maintained through different transmission cycles in nature. Meningoencephalitis due to a S. neurona-like protozoan was reported in marine mammals: Pacific harbor seals (Phoca vitulina richardsi) (Lapointe et al., 1999) and Alaskan sea otters (Enhydra lutris) (Rosonke et al., 1999). The parasitic zoites in infected tissues reacted with S. neurona antibodies although the habitat and behavior of the harbor seals would most likely exclude contact 16 with opossum feces, and it would appear highly unlikely that these marine mammals became infected from the same sporocyst source as the horse. Thus, it is possible that the protozoa described in numerous species, including raccoons (Dubey et al., 1990), northern gannets (Spalding et al., 2002), sheep (Dubey et al., 1989b), skunks (Dubey et al., 1996), and mink (Dubey and Hedstrom, 1993) as “S. neurona-like”, are actually different species with varying life cycles. To date, the population structure and the relationships among S. neurona strains threatening equine health remain ill defined. Further, the question should be raised about how much variation exists between different S. neurona strains that infect horses. These previous studies have found considerable evidence of genetic heterogeneity among Sarcocystis strains from different species of vertebrates, and there is now mounting evidence suggesting that a series of host-adapted genotypes/strains/species of the parasites exists. Unfortunately, recognition of S. neurona as the causative agent of EPM in horses and other mammals is probably an oversimplification. First, there might be some strains containing a mixture of genotypic alleles. Second, studies to date have failed to identify recombinant genotypes, suggesting that most genotypically defined strains are reproductively isolated populations. This contrasts with the present taxonomy of the species, which includes only one recognized genotype. These studies clearly indicate that the current classification of Sarcocystis spp. affected horses and the species level taxonomy of S. neurona does not appear to reflect molecular phylogenetic analyses or biological reality, and this situation is likely to be a limitation in epidemiological investigations. Indeed, current molecular data provide support for some of the earlier taxonomic separations on the basis of host occurrence (Wenvon, 1926). 17 Although the previous studies have confirmed genetic diversity among Sarcocystis strains from horses and opossums, most have been small studies with few cases. Therefore, there is a need for population-level information on the population structure and genetic diversity of S. neurona and non-S. neurona isolates. A larger study encompassing diverse populations is necessary before we can have a more complete understanding of the population structure and molecular epidemiology of S. neurona and other Sarcocystis spp. Further sampling and molecular description will almost certainly show that S. neurona actually comprises several cryptic species. Determining the population structure and understanding the nature and extent of genetic heterogeneity among S. neurona strains is a task of great importance from an economic and equine health standpoint in the US. This information will be valuable for understanding virulence traits and for unraveling the broad evolutionary history of this species. Improved fingerprinting methods are required if fundamental questions about epidemiology and risk factors are to be addressed. 18 REFERENCES Barta JR (1989) Phylogenetic analysis of the class Sporozoea (phylum Apicomplexa Levine, 1970): evidence for the independent evolution of heteroxenous life cycles. J Parasitol 75: 195—206 Bentz BG, Granstrom DE, Stamper S (1997) Seroprevalence of antibodies to Sarcocystis neurona in horses residing in a county of southeastern Pennsylvania. J Am Vet Med Assoc 210:517—518 Bentz BG, Ealey KA, Morrow J, Claypool PL, Saliki JT (2003) Seroprevalence of antibodies to Sarcocystis neurona in equids residing in Oklahoma. J Vet Diagn Invest 6:597—600 Blears MJ, De Grandis SA, Lee H, Trevors J T (1999) Amplified fragment length polymorphism (AF LP): review of the procedure and its applications. J Ind Microbiol Biotechnol 21 :99—1 14 Blears MJ, Pokomy NJ, Carreno RA, Chen S, De Grandis SA, Lee H, Trevors J T (2000) DNA Fingerprinting of Cryptosporidium parvum Strains Using Amplified Fragment Length Polymorphism (AF LP). J Parasitol 86:838—841 Blythe LL, Granstrom DE, Hansen DE, Walker LL, Bartlett J, Stamper S (1997) Seroprevalence of antibodies to Sarcocystis neurona in horses residing in Oregon. J Am Vet Med Assoc 210:525—527 Boch J, Bierschenck A, Erber M, Weiland G (1979) Sarcocystis und T oxoplasma- Infektionen bei Schlachtschafen in Bayem. Berl Munch Tierarztl Wochenschr 92: 137— 141 Box ED, Duszynski DW (1980) Sarcocystis of passerine birds: sexual stages in the opossum (Didelphis virginiana). J Wildl Dis 18:209—215 Cheadle MA, Dame JB, Greiner EC (2001a) Sporocyst size of strains of Sarcocystis shed by the Virginia opossum (Didelphz's virginiana). Vet Parasitol 95:305—311 Cheadle MA, Tanhauser SM, Dame JB, Sellon DC, Hines M, Ginn PE, MacKay RJ, Greiner EC (2001b) The nine-banded armadillo (Dasypus novemcinctus) is an intermediate host for Sarcocystis neurona. Int J Parasitol 312330—335 Cheadle MA, Yowell CA, Sellon DC, Hines M, Ginn PE, Marsh AE, Dame JB, Greiner EC (2001c) The striped skunk (Mephitis mephitis) is an intermediate host for Sarcocystis neurona. Int J Parasitol 31 :843—849 19 Dame JB, MacKay RJ, Yowell CA, Cutler TJ, Marsh A, Greiner EC (1995) Sarcocystis falcatula from passerine and psittacine birds: synonymywith Sarcocystis neurona, agent of equine protozoal myeloencephalitis. J Parasitol 81 :93 0—935 Dubey JP, Davis SW, Speer CA, Bowman DD, de Lahunta A, Granstrom DE, Topper MJ, Hamir AN, Cummings JF, Suter MM (1991) Sarcocystis neurona n. sp. (Protozoa: Apicomplexa), the etiologic agent of equine protozoal myeloencephalitis. J Parasitol 77 :2 1 2—2 1 8 Dubey JP, Hamir AN, Hanlon CA, Topper MJ, Rupprecht CE (1990) Fatal necrotizing encephalitis in a raccoon associated with a Sarcocystis-like protozoan. J Vet Diagn Invest 2:345—347 Dubey JP, Hamir AN, Niezgoda M, Rupprecht CE (1996) A Sarcocystis neurona-like organism associated with encephalitis in a striped skunk (Mephitis mephitis). J Parasitol 82: 1 72—1 74 Dubey JP, Hedstrom OR (1993) Meningoencephalitis in mink associated with a Sarcocystis neurona-like organism. J Vet Diagn Invest 5: 467—471 Dubey JP, Lindsay DS (1998) Isolation in immunodeficient mice of Sarcocystis neurona from opossum (Didelphis virginiana) faeces, and its differentiation fiom Sarcocystis falcatula. Int J Parasitol 28:1823—1828 Dubey JP, Lindsay DS (1999) Sarcocystis speeri n. sp. (Protozoa: Sarcocystidae) from the opossum (Didelphis virginiana). J Parasitol 85:903—909 Dubey JP, Lindsay DS, Saville WJA, Reed SM, Granstrom DE, Speer CA (2001a) A review of Sarcocystis neurona and equine protozoal myeloencephalitis (EPM). Vet Parasitol 95:89—131 Dubey JP, Speer CA, Fayer R (1989a) Sarcocystosis of Animals and Man. CRC Press, Inc, Boca Raton Dubey JP, Speer CA, Munday BL, Lipscomb TP (1989b) Ovine sporozoan encephalomyelitis linked to Sarcocystis infection. Vet Parasitol 34:159—163 Dubey JP, Saville WJ, Stanek JF, Lindsay DS, Rosenthal BM, Oglesbee MJ, Rosypal AC, Njoku CJ, Stich RW, Kwok OC, Shen SK, Hamir AN, Reed SM (2001b) Sarcocystis neurona infections in raccoons (Procyon lotor): evidence for natural infection with sarcocysts, transmission of infection to opossums (Didelphis virginiana), and experimental induction of neurologic disease in raccoons. Vet Parasitol 100:117—129 Dubey JP, Rosenthal BM, Speer CA (2001c) Sarcocystis lindsayi n.sp. (Protozoa: Sarcocystidae) from the South American opossum, Didelphis albiventris from Brazil. I Eukar Microbiol 48:595-603 20 Dubey JP, Saville WJ, Lindsay DS, Stich RW, Stanek JF, Speer CA, Rosenthal BM, Njoku CJ, Kwok OC, Shen SK, Reed SM (2000) Completion of the life cycle of Sarcocystis neurona. J Parasitol 86: 1276—1280 Fayer R, Dubey JP, Leek RG (1982) Infectivity of Sarcocystis spp. from bison, elk, moose and cattle for cattle vis sporocysts from coyotes. J Parasitol 681681—685 Fenger CK, Granstrom De, Langemeier J L, Gajadhar A, Cothran G, Tramontin RR, Stamper S, Dubey JP (1994) Phylogenetic relationship of Sarcocystis neurona to other members of the family Sarcocystidae based on the sequence of the small ribosomal subunit gene. J Parasitol 79:966—975 Fenger C K, Granstrom DE, Langemeier J L, Stamper S, Donahue J M, Patterson J S, Gajadhar AA, Marteniuk JV, Xiaomin Z, Dubey JP (1995) Identification of opossums (Didelphis virginiana) as the putative definitive host of Sarcocystis neurona. J Parasitol 81 :91 6—9 1 9 Fenger CK, Granstrom DE, Gajadhar AA, Williams NM, McCrillis SA, Stamper S, Langemeier J L, Dubey JP (1997) Experimental induction of equine protozoal myeloencephalitis in horses using Sarcocystis sp. sporocysts from the opossum (Didelphis virginiana). Vet Parasitol 68: 199—2 1 3 Granstrom DE, Dubey JP, Davis SW, Payer R, Fox JC, Poonacha KB, Giles RC, Comer PF (1993) Equine protozoal myeloencephalitis: antigen analysis of cultured Sarcocystis neurona merozoites. J Vet Diagn Invest 5:88—90 Granstrom DE, McPherson J M, Gaj adhar AA, Dubey JP, Tramontin R, Stamper S (1994) Differentiation of Sarcocystis neurona from eight related coccidia by random amplified polymorphic DNA assay. Mol Cell Probes 82353—356 Heydron AO, Gestrich R, Mehlhom H, Rommel M (1975) Proposal for a new nomenaclature of the Sarcosporidia. Z Parasienkd 48:73—82 Heydron AO, Matuschka FR (1981) Zur Endwirtspezifitat der vom Hund ubertragenen sarkosporidienarten. Z Parasitenkd 66:231—234 J akes KA (1998) Sarcocystis mucosa in Bennetts wallabies and pademelons from Tasmania. J Wildl Dis 34:594—599 Jenkins MC, Ellis JT, Liddell S, Ryce C, Munday BL, Morrison DA, Dubey JP (1999) The relationship of Hammondia hammondi and Sarcocystis mucosa to other heteroxenous cyst-forming coccidia as inferred by phylogenetic analysis of the 18S SSU ribosomal DNA sequence. Parasitology 119235—142 Lapointe JM, Duignan PJ, Marsh AE, Gulland FM, Barr BC, Naydan DK, King DP, Farman C, Burek KA, Lowenstine J L (1998) Meningoencephalitis due to a Sarcocystis 21 neurona-like protozoan in Pacific Harbor seals (Phoca vitulina richardsi). J Parasitol 84:1184—1189 MacKay RJ (1997) Equine protozoal myeloencephalitis. Vet Clin North Am Equine Pract 1 3 :79—96 MacKay RJ, Granstrom DE, Saville WJ, Reed SM (2000) Equine protozoal myeloencephalitis. Vet Clin North Am Equine Pract 16:405—425 Mansfield LS, SchottII HC, Murphy AJ, Rossano MG, Tanhauser SM, Patterson J S, Nelson K, Ewart SL, Marteniuk JV, Bowman DD, Kaneene J B (2001) Comparison of Sarcocystis neurona strains derived from horse neural tissue. Vet Parasitol 95267—178 Marsh AE, Barr BC, Tell L, Bowman DD, Conrad PA, Ketcherside CC, Green T (1999) Comparison of the internal transcribed spacer, ITS-1, from Sarcocystisfalcatula strains and Sarcocystis neurona. J Parasitol 85:750—757 Marsh AE, Barr BC, Tell L, Koski M, Greiner E, Dame J, Conrad PA (1997a) In vitro cultivation and experimental inoculation of Sarcocystisfalcatula and Sarcocystis neurona merozoites into budgerigars (Melopsittacus undulatus). J Parasitol 85:1189—1192 Marsh AE, Barr BC, Lakritz J, Nordhausen R, Madigan J E, Conrad PA (1997b) Experimental infection of nude mice as a model for Sarcocystis neurona-associated encephalitis. Parasitol Res 83 :706—711 Marsh AE, Johnson PJ, Ramos-Vara J, Johnson GC (2001) Characterization of a Sarcocystis neurona isolate from a Missouri horse with equine protozoal myeloencephalitis. Vet Parasitol 95: 143—154 Mehlhom HW, Hartley WJ, Heydron A0 (1976) A comparative ultrastructural study of the cyst wall of 13 Sarcocystis species. Parasitologica 12:451—467 National Animal Health Monitoring System (NAHMS Equine) (1998) Needs assessment survey results. United States Department of Agriculture: Animal Plant Health Inspection Service: Veterinary Services, Fort Collins, CO, #N207.597 Rosonke BJ, Brown SR, Tomquist SJ, Synder SP, Garner MM, Blythe LL (1999) Encephalomyelitis associated with a Sarcocystis neurona-like organism in a sea otter. J Am Vet Med Assoc 215:1839—1842 Rossano MG, Kaneene JB, Marteniuk JV, Banks BD, Schott HC 2nd, Mansfield LS (2003) A herd-level analysis of risk factors for antibodies to Sarcocystis neurona in Michigan equids. Prev Vet Med 5727—13 Slapeta JR, Modry D, Votypka J, J irku M, Lukes J, Koudela B (2003) Evolutionary relationships among cyst-forming coccidia Sarcocystis spp. (Alveolata: Apicomplexa: 22 Coccidea) in endemic African tree vipers and perspective for evolution of heteroxenous life cycle. Mol Phylogenet Evol 27:464—475 Spalding MG, Yowell CA, Lindsay DS, Greiner EC, Dame JB (2002) Sarcocystis meningoencephalitis in a northern gannet (Moms bassanus). J Wildl Dis 38:432—437 Tanhauser SM, Yowell CA, Cutler TJ, Greiner EC, MacKay RJ, Dame JB (1999) Multiple DNA markers differentiate Sarcocystis neurona and Sarcocystisfalcatula. J Parasitol 85:221—228 Vos P, Hogers R, Bleeker M, Reijans M, van de Lee T, Homes M, Frijters A, Pot J, Peleman J, Kupier M, Zabeau M (1995) AF LP: a new technique for DNA fingerprinting. Nucleic Acids Res 23:4407—4414 Wenvon CM (1926) Protozoology. Bailliere, Tindall and Cox, London 23 CHAPTER 2 PURIFICATION OF SARC 0C YS T IS NE URONA MATERIALS This chapter is based on the following two studies: 1. Purification of Sarcocystis neurona sporocysts from opossum (Didelphis virginiana) using potassium bromide discontinuous density gradient centrifugation. 2. Generally applicable methods to purify intracellular coccidia from cell cultures and to quantify purification efficacy using quantitative PCR. 24 1. Purification of Sarcocystis neurona sporocysts from opossum (Didelphis virginiana) using potassium bromide discontinuous density gradient centrifugation Published as original paper: Hany M. Elsheikha 1, Alice J. Murphy 2, Scott D. Fitzgerald 2, Linda S. Mansfield 2’ 3, Jeffrey P. Massey 4, and Mahdi A. Saeed 1 (2003) Purification of Sarcocystis neurona sporocysts from opossum (Didelphis virginiana) using potassium bromide discontinuous density gradient centrifirgation. Parasitology Research 90: 104—109. 1 Molecular Epidemiology Laboratory, Food Safety and Toxicology Center, Michigan State University, MI 48824, East Lansing, USA. 2 Animal Health Diagnostic Laboratory, Michigan State University, MI 48824, East Lansing, USA. 3 Department of Microbiology and Molecular Genetics, Michigan State University, MI 48824, East Lansing, USA. 4 Molecular Biology Section, Michigan Department of Community Health, MI 48909, Lansing, USA. 25 ABSTRACT This report describes a new, inexpensive procedure for the rapid and efficient purification of Sarcocystis neurona sporocysts from Opossum small intestine. S. neurona sporocysts were purified using a discontinuous potassium bromide density gradient. The procedure provides a source of sporocyst wall and sporozoites required for reliable biochemical characterization and for immunological studies directed at characterizing antigens responsible for immunological responses by the host. The examined isolates were identified as S. neurona using random amplified polymorphic DNA primers and restriction endonuclease digestion assays. This method allows the collection of large numbers of highly purified S. neurona sporocysts without loss of sporocyst viability as indicated by propidium iodide permeability and cell culture infectivity assays. In addition, this technique might also be used for sporocyst purification of other Sarcocystis spp. 26 INTRODUCTION Since the early 19905, the protozoan parasite Sarcocystis neurona has been increasingly recognized as an important pathogen infecting the equine central nervous system, causing equine protozoal myeloencephalitis (EPM) (Fenger et al., 1997; Dubey et al., 2001; Mansfield et al., 2001; Marsh et al., 2001). Sarcocystis spp. are ubiquitous parasites and have an obligatory two-host life cycle, with sexual reproduction and sporocyst formation occurring in the definitive host and schizogony and sarcocyst formation occurring in the intermediate host (Gardiner et al., 1988). Although, the life cycle is not completely understood, natural and experimental transmission studies with S. neurona indicate that the organism lacks host specificity and can be transmitted via the fecal-oral route between opossums, the definitive host (F enger et al., 1995, 1997; Dubey and Lindsay, 1998) and other mammalian intermediate hosts of several different species such as cats, armadillos, skunks, and raccoons (Dubey et al., 2000; Cheadle et al., 2001c, 2001d; Hamir and Dubey, 2001). Horses are exposed to multiple Sarcocystis spp. in the environment that may complicate the understanding of the life cycle of S. neurona, and the pathogenesis and diagnosis of EPM. S. dasypi and S. diminuta sarcocysts have been detected from the tongues of arrnadillos in North America (Lindsay et al., 1996). In addition, at least three species of Sarcocystis (S. neurona, S. falcatula, S. speeri) have been shown to use the North American opossum (Didelphis virginiana) as their definitive host (Fenger et al., 1995; Tanhauser et al., 1999; Cheadle et al., 2001a). These species have the sporocyst stage in the opossum small intestine and are difficult to differentiate based on morphological characteristics alone. 27 Some reports describe sporocyst isolation from opossum small intestine, but none provides pure and sufficient parasite material to carry out genetic, biochemical, and immunological investigations (Murphy and Mansfield, 1999; Cheadle et al., 2001a; Dubey et al., 2001; Ellison et al., 2001). Although sporocysts represent a valuable source of stage specific parasite material, there are limitations concerning their use, including: (1) purification is time-consuming and provides low yield, and (2) there are some Sarcocystis species that produce low numbers of sporocysts and/or shed sporocysts intermittently in the feces during infection. Also, infections with multiple Sarcocystis species may occur. The infected small intestine contains several developmental stages of the parasite. The ability to retrieve intracellular Sarcocystis oocysts/sporocysts stages has been hampered by the lack of a method for separating the parasitized cells (situated exclusively on the brush border layer of the mucosa) from the majority of non-parasitized cells of the intestinal wall. Therefore, there is a need to develop alternatives to current methods to allow not only research into the biology of the Sarcocystis life cycle, but also the characterization of the antigens and the molecular constituents of the sporocyst stage of the parasite. The objectives of the present study were (1) to maximize the sporocyst yield to assure the isolation of a quantitatively representative pure sporocyst population for subsequent studies, (2) to genotype the investigated Sarcocystis spp. sporocysts, (3) to use a dye permeability assay using propidium iodide (P1) to determine whether the permeability of the sporocyst wall to P1 would be changed by exposure to a potassium bromide (Kbr) gradient, and (4) to perform a cell culture infectivity assay to examine the effects of Kbr purification on the sporozoites' viability. 28 MATERIALS AND METHODS Parasite isolates. Two Sarcocystis species isolates used during this study were obtained from naturally infected opossums (D. virginiana) that had been killed by automobiles on the roadways in Michigan and that had been submitted for necropsy to the Wildlife Services Agency of the Michigan Department of Natural Resources, and the Animal Health Diagnostic Laboratory, Michigan State University. Sample collection and sporocyst isolation. Fecal specimens were examined for the presence of Sarcocystis sporocysts using saturated NaCl (360 g 1'], sp. gr. 1.21) fecal floatation. It is helpful to start with fecal specimens that have large numbers of sporocysts, however, some animals may harbor the parasite and not shed sporocysts. Some opossums that have been tested negative for Sarcocystis spp. infection using the fecal floatation test were found to be heavily infected with the parasite in the small intestine by examining smears from mucosal scrapings (unpublished data). Therefore, the opossums' intestines were initially screened for Sarcocystis spp. sporocysts by mucosal scrapings, and samples found to be positive for the presence of sporocysts were processed according to the method described by (Murphy and Mansfield, 1999) with some modifications. A few segments of the small intestine from each animal were removed and washed with 0.01 M phosphate-buffered saline, pH 7.4 (PBS). A scraping of mucosa was observed at 200x magnification (Reichert microscope) to confirm the presence of oocysts and/or sporocysts. Then, the remaining small intestine was flushed with sterile PBS to remove the contents before being slit lengthwise. The intestinal content was collected separately. The mucosa was scraped off using the edge of a glass slide to collect the mucosa with the oocysts and/or sporocysts contained within the brush border. The 29 mucosal scrapings were washed three times with PBS by centrifugation at 500 g for 10 min. The pellet was resuspended in 3 vol. of pepsin-NaCl-HCI (0.65% pepsin w/v, 0.86% NaCl w/v, 1% conc. HCl v/v) and incubated at 37°C for 1.5 h with frequent mixing. The slurry was washed 3 times with PBS as mentioned previously and the pellet stored at 4°C in 3 vol. of Hank's balanced salt solution (HBSS) plus 100 U ml'1 of penicillin, 100 ug ml'lof streptomycin, and 1.25 pg ml‘1 of fungizone until further use. A 1to 3 ml aliquot of the semi-digested mucosa was concentrated by centrifugation for 10 min at 500 g. The pellet was decontaminated by suspending in 10 ml of a 10% (v/v) commercial bleach solution consisting of 1 m1 household bleach (5.25% sodium hypochlorite) and 9 ml sterile deionized distilled water, and incubating on ice for 5 min with frequent agitation. The bleach solution was removed by centrifugation (8 min at 3,000 g) at 4°C and the pellet was washed twice with sterile PBS (pH 7.2). The supernatant was removed and the pellet resuspended in 3 ml of T ris-EDTA buffer (Tris base 50 mM, EDTA 10 mM). Fat aggregates, bacteria, and debris contaminate the sporocysts when gut slurry with a high fat content is used. Since the yield and purity of the recovered sporocysts is highly influenced by fat content, two washing procedures were used to decrease the fat content and to improve the sporocyst yield and purity. In the case of specimens that contain excess fat, the sporocyst suspension was vigorously mixed with an equal volume of ethyl ether. Two layers formed; the upper (ethyl ether) layer contained the lipid material, while the sporocysts remained in the aqueous phase. The upper layer was removed and the sporocysts washed twice by centrifugation at 3,000 g for 10 min, 4°C in Tris-EDTA buffer to remove residual ether. The final samples with low fat content were 30 repeatedly washed by centrifugation at the same conditions in Tris-EDTA buffer until a clear supernatant was obtained. The final pellet obtained in both cases was resuspended in a minimal volume of Tris-EDTA buffer before being applied to the gradient. Discontinuous Kbr density gradient. A discontinuous gradient of Kbr (P9881, Sigma, St. Louis, Mo., USA) was prepared. The gradient consisted of three solutions of 5, 15, and 25% (w/v) Kbr in Tris-EDTA buffer. From 7 to 8 ml each of 25, 15, and 5% ice-cold Kbr solutions was carefully layered from bottom to top into 30 ml Nalgene Centrifuge, Oak Ridge Style 3118 transparent polycarbonate tubes (Sigma). From 3 to 4 ml of each concentrated mucosal homogenate was carefully layered on the top of each gradient. The gradients were centrifuged in a CR/CT4.12 centrifuge equipped with a M4 swinging bucket rotor at 3,000 g for 30 min at 4°C. Afier centrifugation, 3 clear white, distinct band was observed at the interface between the 5% and 15% Kbr solutions (Figure 1). This white band contained S. neurona sporocysts free of debris and bacteria. The cleaned sporocysts (Figure 2B) were almost entirely free of debris (Figure 2A). The band was gently aspirated at the interface using a sterile Pasteur pipette, and the recovered sporocysts were diluted threefold with sterile 1>< PBS (pH 7.2) and centrifuged at 3,000 g for 10 min at 4°C. The pellet was resuspended and washed twice using the same buffer and centrifirgation conditions. The final recovered pellet was resuspended in sterile l>< PBS. The yield of sporocysts was determined microscopically by counting the number of sporocysts using a hemocytometer. 31 Figure 1. Centrifuge tubes showing a white band (arrows) of Sarcocystis neurona sporocysts afier potassium bromide discontinuous gradient centrifugation. (A) At 3,000 g, 30 min, and 4°C. (B) At 16,000 g, 1 h, and 4°C. Images in this dissertation are presented in color. 32 Finally, purified sporocysts were resuspended in a minimal volume of storage media 6. g. HBSS containing an antibiotic-antimycotic mixture, as mentioned previously, at 4°C. Each lot of purified sporocysts was tested by the dye permeability assay after purification as described below. In some cases, higher centrifugal forces were needed to enhance the purity of the final suspension. Here, the Kbr gradient was centrifuged under conditions described for cesium chloride gradient purification of Cryptosporidium parvum oocysts (Taghi-Kilani and Sekla, 1987), 16,000 g for 1 h at 4°C using a Sorvall RC-SC superspeed, refrigerated centrifuge equipped with a Sorvall SM-24 rotor. Morphometric characterization. Sporocyst type was elucidated using morphological analysis (Cheadle et al., 2001a). An aliquot of 20 ul of purified sporocysts from each isolate was quickly mixed with two warm drops of melted agar (2 mg BACTO-agar powder (Difco)/ 100 ul 0.2 M Tris-HCl, pH 8.0). The sporocyst agar suspension was placed on the center of a clean glass slide and covered with a clean glass coverslip. The slide was then observed using light microscopy (Reichert) for morphological characteristics. Then, immersion oil with 1000X magnification was employed for the measurement of the sporocysts. A laser scanning microscope (LSM Zeiss 210) with 488 nm laser line was used to take images (Figure 2). At least 25 sporocysts per isolate were measured using an ocular micrometer calibrated against its stage micrometer slide after the agar had hardened. Minimum and maximum values are given, followed in parentheses by the arithmetic mean and standard deviation. All measurements were in micrometers. Genotypic identification. The genotype of Sarcocystis species sporocysts was identified according to the method described by Tanhauser et al., (1999). Genotyping of 33 the isolate was carried out using DNA extracted from purified sporocysts from opossums using the Dneasy Tissue kit (Qiagen). Pure sporocysts were subjected to freezing in liquid nitrogen followed by thawing for six cycles in order to break the oocyst and sporocyst wall. Then, the sporozoite DNA was extracted according to the manufacturer's instructions. Genotypic analysis of the DNA from Sarcocystis species isolates was performed using PCR analysis with primers JNBZS/JD396 (Tanhauser et al., 1999). The primers were used to amplify a 334 bp amplicon from the sporocyst DNA of the isolates. The amplified 334 bp PCR products were subsequently digested with HindIII or Hinfl restriction enzymes to differentiate bands characteristic of S. neurona and S. falcatula. S. speeri does not amplify with these primers (S. Tanhauser, personal communication). The digested DNA products were subjected to electrophoretic separation on a 2% agarose gel with appropriate size markers, stained with ethidium bromide, visualized under UV light and compared with an S. neurona (MIHl) isolate cultivated from horse neural tissue and an S. falcatula (Cornell) isolate obtained from sporocysts shed from an opossum fed a wild-caught grackle (Quiscalus sp.). Dye permeability assay. The PI exclusion assay was used to indicate sporocyst viability and potential infectivity. The dye permeability assay was used as described previously (Campbell et al., 1992). This assay was initially used to assess the viability of C. parvum oocysts. We have adapted this assay to evaluate the viability of Sarcocystis spp. sporocysts. A stock solution of PI (Sigma, P4170) 1 mng'1 in 0.1 M PBS, pH 7.2 was added to aliquots of the purified sporocysts suspended in 1.5 ml microcentrifuge tubes at a ratio of 1:10 (v/v). The contents of the tubes were mixed gently and the tubes were incubated in the dark at 37°C for 2 to 3 h. Each sample was examined by both 34 epifluorescence and differential interference contrast (DIC) microscopy. At least 50 sporocysts in each sample were examined and scored for viability. Cell culture infectivity assay. Maintenance, passaging, and growth of equine dermal cells (ATCC, CCL-57) (American Type Culture Collection, Manassas, Va.) were performed as described by Mansfield et al., (2001). For infectivity assays, cells were grown on sterile 12-mm2 coverslips positioned at the bottom of 24-well tissue culture plate in a 5% C02 atmosphere at 37°C to approximately 70-90% confluence (24-48 h). Cultures were infected with sporozoites obtained from sporocysts before and after purification as described by Murphy and Mansfield, (1999). Cultures were placed in a 5% C02 incubator at 37°C. During in vitro cultivation, the medium was changed every i 3 days to maintain the pH within the range 7.2-7 .6. Both light inverted and epifluorescence microscopy were used to assess plaques in the infected cell monolayers. After incubation for 4 weeks, the medium was aspirated and the infected monolayers were washed once with 1X PBS and coverslips were fixed in 10% phosphate-buffered formalin for 30 min, placed in 100% methanol (room temperature) for 10 min, and stained with Giemsa. Coverslips were attached to glass microscope slides with Perrnount (Fisher Scientific, Fair Lawn, NJ, USA) and examined by light microscopy to observe the sporozoite and schizont stages of the life cycle. 35 RESULTS We developed a protocol for collecting the S. neurona parasites from the brush border of the intestinal mucosa, thereby increasing the ratio of parasites to intestinal cells and reducing other debris and contaminants. By centrifugation on a discontinuous Kbr gradient, it was possible to recover and purify Sarcocystis sporocysts from the intestinal mucosa. The use of commercial bleach for 5 min prior to gradient centrifugation greatly facilitated the separation of oocysts and sporocysts from intestinal cells and fecal debris. We routinely isolated 10-20><106 sporocysts in 3-4 h without the use of specialized or expensive equipment. Oocysts were ovoid to subspherical, smooth walled, colorless, and lacked oocyst residuum, polar granule or micropyle. The sporocysts were ovoid to round, 9.2-11.3X6.0—8.1 (10.3:t0.55><7.16:t0.56) with length to width ratios of 1.3-1.6. The sporocyst wall was colorless, thin and smooth. All sporocysts examined possessed four sporozoites and a distinct residuum (Figure 2C). Figure 2. Photomicrographs of cleaned and crude Sarcocystis neurona sporocysts. Crude sporocysts were collected and cleaned from intestinal mucosa of opossum. (A, B) Phase contrast microscopy of crude and cleaned sporocysts. Scale bar = 25 pl. (C) Oblique illumination of purified sporocysts. Note the residual body (white arrow), sporozoite (black arrow), oocyst wall (black arrowhead), and sporocyst wall (white arrowhead). Scale bar = 100 pl. 36 The total mean recovery of sporocysts after applying them to the gradient was 86.89ziz4.1% and the recovered sporocysts were greater than 98% pure (Table 1). We compared the current method (mucosal scraping) for isolation of Sarcocystis spp. sporocysts from opossum small intestine to the Kbr-gradient centrifugation method and found that the purity and viability of the isolated sporocysts as well as the freedom of bacterial contamination were improved (Table 1). Table 1. Comparison between mucosal scraping (current method) and potassium bromide discontinuous gradient centrifugation method for the isolation of Sarcocystis spp. sporocysts from opossum small intestine. . . 'Current method (mucosal {Potassium bromidediscontrnuousgradient Criteria ; . i . , scraping) gmethod {Purity (%) F40 ,, -_ l>98 H - Viability - i - . 2t . (%) 192 3 0 6 g98 68 107 {Bacteria lTNTC immutable o The relative purity was assessed microscopically by counting the number of sporocysts and all other particles or debris of any size/microscopic field (averaged over ten randomly selected fields). This procedure was performed on each specimen before and after purification. In both cases, the ratio of sporocysts versus debris was compared to quantify purity percentage before and after Kbr-gradient purification. o The viability of sporocysts was determined using the propidium iodide exclusion assay. 0 TNTC = too numerous to count. 37 1 L.....___. 4M.L~-V_WL.-.;L_-L- _ . RFLP analysis showed banding patterns characteristics of only S. neurona for these two opossum isolates. They both had bands at 180, and 154 bp with Hinfl digestion, and a single band at 334 bp with HindIII digestion (Figure 3). MIOP MIHl Cornell M PHer PHfHd PHfHd Figure 3. Results of random amplified polymorphic patterns using the primers JNB25/JD396 and restriction enzyme digestion with Hinfl and HindIII of sporocysts DNA preparations from Sarcocystis neurona isolates from Michigan opossums (MIOP), S. neurona (MIHl) isolate from horse, and S. falcatula (Cornell) isolate from opossrun. M 100-bp ladder marker; P PCR product; Hf Hinfl digest; Hd Hindlll digest. Note, afier HindIII digestion of PCR 334 bp amplicons, S. neurona amplicon is unaffected while S. falcatula amplicon yields bands at 164, 108, and 62 bp. After Hinfl digestion, S. neurona amplicon yields bands at 180 and 154 bp while S. falcatula amplicon yields bands at 170 and 164 bp. The viability of the purified sporocysts was estimated using propidium iodide dye exclusion assay and was found to be 98.68i1.07%. Successful cultivation of the two S. neurona isolates was observed from sporozoites obtained from the crude and the clean sporocysts, before and after gradient purification, respectively. The in vitro infectivity assays showed that S. neurona sporozoites obtained from Kbr-purified sporocysts were viable and maintained their infectivity afier purification. Generally, cultures infected with sporozoites obtained from purified sporocysts appeared morphologically healthier than those infected with the sporozoites obtained from crude sporocysts. 38 DISCUSSION The goal of these experiments was to develop a reliable method for the purification of large quantities of viable S. neurona or any other Sarcocystis spp. sporocysts. The infected opossum gut contains populations of oocyst and sporocyst stages of the parasite, but its use for obtaining purified material of these intracellular Sarcocystis stages has been hampered by the lack of a method for separating the parasitized cells (situated exclusively on the brush border layer of the mucosa) from the majority of the non-parasitized cells of the intestinal wall. In the present study, the potassium bromide discontinuous gradient was used to isolate pure S. neurona oocysts/sporocysts from the intestinal mucosa of opossums (D. virginiana). Potassium bromide discontinuous i gradient has been effectively used for the purification of C. parvum oocysts (Entrala et aL,2000) Under natural conditions, the oocyst wall is so thin that it often ruptures mechanically. Free sporocysts, released into the intestinal lumen, are passed in feces (Dubey et al., 1989). The recovery of many intact oocysts post-purification that still have wall and enclose two sporocysts (Figure 2C) may result in increased viability of the purified sporocysts due to the protection afford by the cell wall. Based on the measured dimensions of the sporocysts, the Sarcocystis spp. isolates under investigation appeared most closely related to the S. neurona strain (Cheadle et al., 2001a). However, there were no unique internal morphological characteristics of the isolates. The use of agar overlay helped to hold the sporocysts in position, which in turn prevented deformity of the sporocysts and preserved their dimensional structure. 39 For in vivo and in vitro studies of S. neurona, it is important to use well-defined isolates of S. neurona because at least five strains with structurally similar sporocysts have been described from Opossum isolates. Molecular characterization of Sarcocystis sporocysts collected from opossum has been shown to be practical compared to the time- consuming and expensive use of mice to type Sarcocystis isolates (Cheadle et al., 2001b). In the present study, PCR-RF LP analysis identified the two opossum isolates as S. neurona based on characteristics of their banding patterns. However, some opossums have more than one Sarcocystis spp. This purification method should provide the basis for future experiments to develop clonal isolates from individual oocysts or sporocysts. This method improved the viability of the purified samples. PI is a fluorogenic vital dye that indicates the presence of live organism. PI exclusion from the nucleus is based on membrane integrity, and organisms are considered dead (non-viable) when PI is included in the nucleus (Griffiths et al., 1994). Kbr is a hyperosmotic compound, and the recovery of sporocysts is affected by sporocyst viability, since only water impermeable sporocysts can be recovered. The viability of the Kbr-purified S. neurona sporocysts was remarkably high. This enrichment concentration of viable sporocysts has been previously reported in C. parvum oocysts, when hyperosmotic sucrose density or zinc sulfate flotation techniques were used for oocyst purification (Bukhari and Smith, 1995). Infectivity is an important issue that is currently addressed mainly by using animal models. Cell culture seems to be a more cost-effective, reproducible and practical approach. S. neurona can be cultivated in equine dermal cells for 6 months with no apparent loss of viability. S. falcatula grow briefly and die in 4-6 weeks (L.S. Mansfield and A]. Murphy, unpublished data). The sporozoites obtained from purified sporocysts 40 of the two parasite isolates continued to grow in equine dermal cells for at least 2 months, which supports the notion that the genotype of the parasite isolates is S. neurona. This technique is easy to perform, convenient, and allows the rapid purification of S. neurona sporocysts which will in turn enhance research into the biology of host cell- parasite interactions, and enable amplification of parasite materials for further immunological, biochemical, and molecular studies. Axenic sporocysts, suitable for in vitro culture, can be obtained using this method. Therefore, it could also be used for the assessment of the viability of S. neurona sporocysts isolated fi'om opossums. This technique is a general method that can be applied for the purification of sporocysts of any other Sarcocystis species from opossum small intestine. 41 REFERENCES Bukhari Z, Smith HV (1995) Effects of three concentration techniques on viability of Cryptosporidium parvum oocysts recovered from bovine feces. J Clin Microbiol 33:2592—2595 Campbell AT, Robertson LJ, Smith HV (1992) Viability of Cryptosporidium parvum oocysts: correlation of in vitro excystation with inclusion or exclusion of fluorogenic vital dyes. Appl Environ Microbiol 58:3488—3493 Cheadle MA, Dame JB, Greiner EC (2001a) Sporocyst size of isolates of Sarcocystis shed by the Virginia opossum (Didelphis virginiana). Vet Parasitol 95:305—311 Cheadle MA, Tanhauser SM, Scase TJ, Dame JB, MacKay RJ, Ginn PE, Greiner EC (2001b) Viability of Sarcocystis neurona sporocysts and dose titration in gamma- interferon knockout mice. Vet Parasitol 95:223—231 Cheadle MA, Tanhauser SM, Dame J B, Sellon DC, Hines M, Ginn PE, MacKay RJ, Greiner EC (2001c) The nine-banded armadillo (Dasypus novemcinctus) is an intermediate host for Sarcocystis neurona. Int J Parasitol 31 :330—335 Cheadle MA, Yowell CA, Sellon DC, Hines M, Ginn PE, Marsh AE, Dame JB, Greiner EC (2001d) The striped skunk (Mephitis mephitis) is an intermediate host for Sarcocystis neurona. Int J Parasitol 31 :843—849 Dubey JP, Lindsay DS (1998) Isolation in immunodefecient mice of Sarcocystis neurona fiom opossum (Didelphis virginana) faeces, and its differentiation from Sarcocystis falcatula. Int J Parasitol 28:1823—1828 Dubey JP, Speer CA, F ayer R (1989) General biology: In: Sarcocystosis of animal and man. CRC Press, Boca Raton, pp 2—13 Dubey JP, Saville WJ A, Lindsay DS, Stich RW, Stanek J F, Speer CA, Rosenthal BM, Njoku CJ, Kwok OCH, Shen SK, Reed SM (2000) Completion of the life cycle of Sarcocystis neurona. J Parasitol 86:1276—1280 Dubey JP, Mattson DE, Speer CA, Hamir AN, Lindsay DS, Rosenthal BM, Kwok OCH, Baker RJ, Mulrooney DM, Tomquist SJ, Gerros TC (2001) Characteristics of a recent isolate of Sarcocystis neurona (SN7) from a horse and loss of pathogenicity of isolates SN6 and SN7 by passages in cell culture. Vet Parasitol 95:155-166 Ellison SP, Greiner E, Dame JB (2001) In vitro culture and synchronous release of Sarcocystis neurona merozoites from host cells. Vet Parasitol 95:251—261 Entrala E, Molina-Molina J, Rosales-Lombardo M, Sanchez-Moreno M, Mascaro- Lazcano C (2000) Cryptosporidium parvum: oocysts purification using potassium 42 bromide discontinuous gradient. Vet Parasitol 92:223—226 F enger CK, Granstrom DE, Langemeier J L, Stamper S, Donahue J M, Patterson J S, Gajadher AA, Marteniuk JV, Xiaomin Z, Dubey JP (1995) Identification of opossums (Didelphis virginiana) as the putative definitive host of Sarcocystis neurona. J Parasitol 81 :916—919 Fenger CK, Granstrom DE, Gajadher AA, Williams NM, McCrillis SA, Stamper S, Langemeier J L, Dubey JP (1997) Experimental induction of equine protozoal myeloencephalitis in horses using Sarcocystis sp. sporocysts from the opossum (Didelphis virginiana). Vet Parasitol 68:199—213 Gardiner CH, Fayer R, Dubey JP (1988) An atlas of protozoan parasites in animal tissues. Agriculture Handbook, vol 651. USDA, Agricultural Research Service, pp 40—45 Griffiths JK, Moore R, Dooley S, Keusch GT, Tzipori S (1994) Cryptosporidium parvum infection of Caco-2 cell monolayers induces an apical monolayer defect, selectively increases transmonolayer permeability, and causes epithelial cell death. Infect Immun 62:45 06—45 14 Hamir AN, Dubey JP (2001) Myocarditis and encephalitis associated with Sarcocystis neurona infection in raccoons (Procyon lotor). Vet Parasitol 95:335—340 Lindsay DS, McKown R, Upton SJ, McAllister CT, Toivio-Kinnucan MA, Veatch J K, Blagbum BL (1996) Prevalence and identity of Sarcocystis infections in armadillos (Dasypus novemcinctus). J Parasitol 82:518—520 Mansfield LS, Schott HC, Murphy AJ, Rossano MG, Tanhauser SM, Patterson J S, Nelson K, Ewart SL, Marteniuk JV, Bowman D, Kaneene JB (2001) Comparison of Sarcocystis neurona isolates derived from horse neural tissue. Vet Parasitol 952167—178 Marsh AE, Johnson PJ, Ramos-Vara J, Johnson GC (2001) Characterization of a Sarcocystis neurona isolate from a Missouri horse with equine protozoal myeloencephalitis. Vet Parasitol 952143—154 Murphy AJ, Mansfield LS (1999) Simplified technique for isolation, excystation, and culture of Sarcocystis species from opossums. J Parasitol 85:979—981 Taghi-Kilani R, Sekla L (1987) Purification of Cryptosporidium oocysts and sporozoites by cesium chloride and Percoll gradients. Am J Trop Med Hyg 36:505—508 Tanhauser SM, Yowell CA, Cutler TJ, Greiner EC, MacKay RJ, Dame JB (1999) Multiple DNA markers differentiate Sarcocystis neurona and Sarcocystisfalcatula. J Parasitol 85:221—228 43 2. Generally applicable methods to purify intracellular coccidia from cell cultures and to quantify purification efficacy using quantitative PCR Submitted for publication as original paper: HM. Elsheikha], EM. Rosentha12,A.J. Murphyj, D.B. Dunamsz, D. A. Neelis’, and LS. Mansfield“4 (2004) Generally applicable methods to purify intracellular coccidia from cell cultures and to quantify purification efficacy using quantitative PCR. Journal of Microbiological Methods. 1 Department of Large Animal Clinical Sciences, College of Veterinary Medicine, Michigan State University, East Lansing, MI 48824, USA. 2 Animal Parasite Diseases Laboratory, Agricultural Research Service, United States Department of Agriculture, Beltsville Agricultural Research Center, Animal and Natural Resources Institute, Building 1080, BARC-East, Beltsville, MD 20705, USA. 3 Diagnostic Center for Population and Animal Health, College of Veterinary Medicine, Michigan State University, East Lansing, MI 48824, USA. 4 Department of Microbiology and Molecular Genetics, College of Veterinary Medicine, Michigan State University, East Lansing, MI 48824, USA. 44 ABSTRACT The objective of this study was to evaluate the utility of a simple, efficient, and rapid method for the isolation of Sarcocystis neurona merozoites and Besnoitia darlingi tachyzoites from cultured cells. The efficacy of this purification method was assessed by microscopy, SDS-PAGE, Western Blotting, immune-fluorescence, and three novel quantitative PCR assays. Culture medium containing host cell debris and parasites was eluted through PD-lO desalting columns. This purification method was compared to alternatives employing filtration through a cellulose filter pad or filter paper. The estimated recovery of S. neurona merozoites purified by the coltunn method was 82% (i: 3.7) of the original merozoites with 97.5% purity. In contrast, estimated recovery of S. neurona merozoites purified by filter pad and filter paper was 40% and 30% with 76% and 83% purity, respectively. The same procedures were applied to purify B. darlingi tachyzoites from cultured cells. Of the original cultured B. darlingi tachyzoites, 94% (i 2.5) were recovered from the PD-lO column with 96.5%, purity whereas percentage recovery of B. darlingi tachyzoites purified by filter pad and filter paper were 51% and 35% with 84% and 88% purity, respectively. All described methods maintained sterility so that purified parasites could be subsequently cultured in vitro. However, purification using a PD-lO column minimized parasite loss and the loss of viability as determined by the trypan blue dye exclusion assay, the rate of parasite production, and plaque forming efficiency in tissue culture. Moreover, column-purified parasites improved the sensitivity of an immune-fluorescent (IFA) analysis and real-time quantitative PCR assays targeted to parasite 18S ribosomal DNA and hsp70 genes. This technique appears generally applicable for purifying coccidia grown in tissue cultures. 45 INTRODUCTION Sarcocystis neurona is the primary causative agent of equine protozoal myeloencephalitis (EPM) in the Americas (Dubey et al., 2001b). S. neurona was first described in 1991 after its isolation from a horse (Dubey et al., 1991). It belongs to the phylum Apicomplexa, a large group of mainly intracellular parasites. Besnoitia darlingi is an Apicomplexan parasite employing opossum (Didelphis virginiana) as an intermediate host and cat as a definitive host (Dubey et al., 2002). Equine dermal cells support the development of S. neurona in cell culture (Mansfield et al., 2001). In addition, bovine monocytes, equine kidney cells, cardiopulmonary endothelial cells, deer testes, vero cells, Afiican green monkey (Cercopithecus aethiops) kidney cells (CV-1), rat myoblasts, and bovine turbinate (BT) cells have been employed successfully to support S. neurona replication (Dubey et al., 2001b). Bovine monocytes, CV-l cells, and BT cells also have been used to cultivate and maintain B. darlingi tachyzoites (Dubey et al., 2002; Elsheikha et al., 2004). One of the major obstacles to research with S. neurona and other coccidia is the difficulty of obtaining abundant parasites free of host cell contaminants. Such pure parasite suspensions would facilitate studies of host-parasite interaction and are needed in order to extract protein and genetic materials for biochemical, immunological, and molecular analyses. Difficulty in obtaining sufficient numbers of S. neurona merozoites free of contaminating host cells or cell debris is generally recognized. Approaches have been reported for separating the parasite from host cells. Dubey et al., (2001a) reported the isolation of merozoites from cell culture by filtration through a syringe filter. However, the syringe filters tend to clog easily. Lindsay and Dubey, (2001) passed the 46 merozoite suspension through a 27-gauge needle attached to a 10 ml syringe in order to rupture host cells, and then passed the suspension through a sterile 3pm filter to remove cellular debris. However, the parasite yield is low using this method. In another study, a calcium ionophore (A23187) was used for synchronous release of parasites from host cells (Dame et al., 2001). This method was slow, complicated for large-scale preparation and resulted in preparations that contained appreciable host cell contamination. Although useful, none of these approaches are well suited for producing an abundance of pure merozoites desirable for immunological, biochemical, and genetic analyses. Quantitative real-time PCR provides a sensitive, accurate and fast method for estimating the abundance of DNA templates by monitoring the accumulation of fluorescent PCR products. In assays that approach 100% efficiency, in which a doubling of product occurs with each successive amplification cycle during the exponential phase, the number of cycles required to significantly exceed background fluorescence (Ct) is inversely proportional to the initial template concentration (Klein et al., 2002; Ginzinger, 2002) Herein we describe the adaptation of a previously described approach using a PD- 10 desalting column (Hemphill et al., 1996) to isolate viable over 80% of the S. neurona merozoites and over 90% of the B. darlingi tachyzoites cultivated in bovine turbinate cells at a purity of over 95%, and describe new real-time quantitative PCR assays to estimate the relative and absolute concentrations of parasite and host DNA in such samples. The described method compares favorably to known alternatives in parasite recovery, purity, and viability. Additionally, the procedure improves the sensitivity of immune—fluorescent staining and of diagnostic PCR amplifications. 47 MATERIALS AND METHODS Parasite strains and tissue cultures. Sarcocystis neurona strains MIHl and SN3 were maintained in bovine turbinate (BT) cell monolayers. BT cells were originally purchased from American Type Culture Collection (ATCC, CRL-1390, Manassas, VA, USA). BT cells were used between passages 15-17 and maintained in 8-10 m1 of complete culture medium using 25-cm2 plastic tissue culture flasks. Cells were fed Eagle's minimum essential medium (EMEM) with Earle's salt supplemented with L- glutamine, 10,000 U ml‘l penicillin G sodium, 10,000 pg ml'1 dihydrostreptomycin, 250 pg ml'1 of amphotericin B, nonessential amino acids, 100 mM sodium pyruvate, and 5 to 10% heat inactivated fetal bovine serum (F BS) (Gibco-Invitrogen, Grand Island, NY, USA). Cells were incubated at 37 °C in 5 % COz/95% air. The B. darlingi MIBDl strain was initially obtained from cysts isolated from a naturally infected opossum (D. virginiana) from Michigan in 2002. Methods used for isolation and maintenance of this strain were previously described (Elsheikha et al., 2004). Stock cell cultures were also maintained by lifting the monolayers with trypsin-EDTA (0.25% trypsin; 1 mM EDTA.4Na, Gibco) once a week and transfening the cells to new culture flasks. Approximately 5X 1040f S. neurona_merozoites and 3 x 104 B. darlingi tachyzoites were used to inoculate new flasks of BT cells. The development of mature schizonts occurred in 3 to 5 days in both parasites after inoculation of the flasks. Purification of S. neurona merozoites. Sarcocystis neurona merozoites were harvested from their feeder cell cultures when about 60—80% of the BT host cells were lysed. Free merozoites were removed from the tissue culture flasks by collecting the medium supernatant. A large portion of the remaining parasites were released from their 48 host cells by gently shaking the side of the flasks 2 to 3 times. A few milliliters of EMEM was added and the merozoites were collected. The preparation containing merozoites and host cell debris was mixed thoroughly, divided into three equal aliquots, and submitted to purification procedures by PD-lO column, cellulose filter pad, or filter paper. All the purification procedures were performed inside a biological safety cabinet under sterile conditions. Purification using a PD-10 column. The merozoite suspension was washed twice in a buffer solution by centrifugation at 1,500g for 5 min at 4°C. Buffer solutions used were either sterile 1x PBS or 1x Hanks’ balanced salt solution (HBSS). The final pellet was resuspended in 2.5 ml of buffer solution and applied to a PD—10 column filled with sephadex G-25M (Amersham Biosciences, Piscataway, NJ, USA), previously equilibrated with approximately 25 ml of the same buffer. The parasites were eluted with 3.5 ml of the same buffer used in equilibration of the column. The eluted, purified merozoites were used for determination of recovery, purity, and viability as well as for evaluating the utility of parasite materials purified by this method for molecular analyses. Purification using cellulose filter pad and filter paper. The cellulose filter pad method for purifying the merozoite suspension required a glass filter assembly GVIillipore, Bedford, Massachusetts, USA) (Figure l). A 47-min glass holder apparatus was designed to handle large volumes of liquids. An autoclavable cellulose filter pad (cat. no. AP1004700, Millipore) was initially rinsed with sterile 1x PBS, pH 7.2 or EMEM immediately prior to use. Then, the merozoite suspension was passed through the filter pad and the filtrate was collected. The filtrate was centrifuged at 1,500g for 5 min at 4°C. The pellet was resuspended in 3.5 ml of 1x PBS. The filter paper method 49 was performed identically using the same conditions, but by using Whatrnan filter paper (cat. no. 1001125, VWR Intemational,West Chester, PA, USA) instead of a filter pad. The purified merozoites in both cases were used for determination of recovery, purity, and viability. Figure 1. Illustration of the filter assembly. It consisted of a removable 250-ml borosilicate glass funnel (l) and support base (2). An anodized aluminum spring clamp (3) sandwiched a cellulose filter pad or filter paper (4) between funnel and support base. Connection to 500-ml Erlenmeyer filter flask (5) is made with a silicone stopper (6). Parasite suspension with debris (7) and relatively pure parasite suspension (8). 50 Determination of merozoites’ yield and purity. In all experiments, the yield of merozoites was determined microscopically by counting individual intact merozoites before and after purification using the four outside squares and one central square of a standard hemocytometer chamber (Sigma, Saint Louis, MO, USA) at 200X magnification on a light microscope (Reichert-Jung, Pegasus Scientific, Frederick, MD, USA). The total number of merozoites/m1 was estimated by averaging the counted squares and corrected for dilution and volume using the formula (Total merozoites counted in 5 squares/5) x 104 = merozoites/ml. Tubes containing the purified merozoites in 3.5-ml aliquots were vortexed for 10 seconds before counting. The purity of the final merozoite preparation was evaluated microscopically by differential counting of intact motile merozoites versus any other recognizable particles or cellular debris. Percentage purity was quantified by comparing the number of merozoites to the number of debris particles before and after purification. Assessment of merozoite viability. The viability of the purified merozoites was assessed by two means (1) trypan blue dye exclusion assay and (2) by evaluating the rate of S. neurona merozoite production and plaque forming efficiency in BT cell culture. Equal numbers (~ 12 x 105) of crude merozoites and merozoites purified by PD-lO column, filter pad, or filter paper were inoculated onto T-25 flasks containing a monolayer of BT cells in complete EMEM. The cultures were incubated at 37°C in 5% C02. T25 tissue culture flasks were mock-inoculated with complete EMEM and served as negative controls. Media was changed 36 h post-inoculation to remove extracellular merozoites and then at 3 to 4 day intervals thereafter. The inoculated cultures were monitored daily using a light inverted microscope (Carl Zeiss, Opton, Columbia, MD, 51 USA) to follow the development of plaques and production of merozoites. The numbers of S. neurona plaques (infected foci) were estimated 10 days post infection (DPI) using an inverted microscope, by counting the number of visible plaques per microscopic field at X200 magnification. These counts were expressed as a mean (i SD) over 10 randomly selected fields. After 13-day culture period, merozoites were harvested from all flasks by rinsing the monolayer twice with medium and counting the number of merozoites using a hemocytometer as described. SDS-polyacrylamide gel electrophoresis (PAGE) and immunoblotting. Proteins were extracted from an equal number of purified and un-purified merozoites by suspending them in a lysing buffer containing 0.5 M Tris (pH 7.4) with 10% sodium dodecyl sulfate (SDS), 20% glycerol and 5% E-mercaptoethanol and heating the mixture at 95°C for 5 min. Solubilized proteins extracted from approximately 107 pure merozoites were loaded onto a PAGE gel in 2, 5, and 10pl aliquots. For comparison, a 15 p1 of solubilized protein extracted from approximately 107 crude merozoites was also loaded onto a lane of the same gel. The gel and a biotinylated molecular weight standard (Bio-Rad, Hercules, CA) were separated electrophoretically on a 12—18% SDS—PAGE gradient minigel at 150 V and transferred to Westran PVDF membrane (Bio-Rad). The remainder of the immunoblotting procedure was performed as described previously (Mansfield et al., 2001). Also, the protein extracts were visualized by silver staining as follow; after SDS-PAGE, gels with the separated proteins were silver stained by fixation for 30 min in 50% methanol-10% acetic acid followed by staining using the Silver Stain Plus kit (Bio-Rad), and scanned using a Microtek ScanMaker III (Microtek Lab). Protein 52 bands from each lane were compared to size standards and patterns of each lane were compared. Immunofluorescence assay (IFA). To prepare IF A slides, serial dilutions of the pure and un-pure merozoite stocks were placed in duplicate in wells of a 12 well slides (20pl/well), air dried and fixed in acetone. Cerebrospinal fluid (CSF) from a horse positive for S. neurona antibodies was placed on each well and incubated for 30 min at 37°C. The slide was washed several times with PBS and then incubated with Fluorescein isothiocyanate- (FITC) - labeled goat anti-horse IgG- h+1 secondary antibodies in PBS and 0.1% Evans blue and incubated at 37°C for an additional 30 min. The slide was washed with PBS, cover-slipped and the optimum dilution of merozoites for IF A slides determined. The stock merozoite solution was diluted with PBS to the optimum and multiple slides prepared through the fixation step and stored frozen at -20°C. Positive sera were applied at 1:2500 and 1:10,000 dilutions in PBS and incubated for 30 min at 37°C followed by washing as above. Serum from a knockout mouse previously infected with S. neurona_was used as a positive control. FITC-labeled goat-anti-mouse IgG-h+1 secondary antibody (Kirkegaard & Perry Labs, Inc.) diluted 1:10 in PBS and 0.1% Evans blue dye was applied. Stained merozoites were observed by epifluorescent microscopy using the SPOT RT slider “F” mount camera (model No. 2.3.1, Diagnostic Instruments Inc) and SPOT RT software V3.3. Diagnostic sensitivity of conventional PCR for S. neurona. Crude merozoites and merozoites purified using a PD-lO column were diluted and adjusted to equal concentations of 1000 merozoites/200 pl 1x PBS. Ten-fold serial dilutions (1000, 100, and 10 merozoites) from both samples were prepared to a final volume of 200p1 in 1x 53 PBS for DNA extraction using the QIAamp® DNA Blood Mini Kit (Qiagen). The Qiagen® Blood and Body Fluid Spin Protocol was followed according to the manufacturer’s recommendations to obtain a high DNA concentration from each dilution. DNA concentration was assessed using a NanoDrop DN-1000 spectrophotometer and NanoDrop software 2.5.4. (N anoDrop Technologies, Rockland, DE). The DNA extracts were subjected to PCR using primers JNB25 and JD396 (Tanhauser et al., 1999). PCR products were electrOphoresed through a 1.8% agarose gel for detection of a 334 bp band, indicative of S. neurona. Purification of B. darlingi tachyzoites. Besnoitia darlingi tachyzoites were purified and analyzed from cultured BT cells using procedures identical to those described for S. neurona merozoites. For each purification method used, recovery, purity, and viability of B. darlingi tachyzoites were measured using methods identical to those used for S. neurona merozoites. Quantitative PCR assay design. We developed a new quantitative PCR assay employing specific amplification of a small portion of nuclear 188 rDNA using primers complementary to sequences conserved among coccidia but distinct from other taxa. Additional selection criteria included minimal non-specific annealing, low GC content, balanced primer length, and an amplification product which required fewer than 20 cycles in order to exceed threshold fluorescence (Ct) and dissociated over a narrow and reproducible temperature range. After evaluating several candidate primer combinations, we chose one pair that best met these criteria. Similar criteria were used in developing quantitative assays for Besnoitia darlingi based on the Heat Shock Protein 70 gene (hsp70) and for vertebrate 18S rDNA (Table 1). 54 Table 1. qPCR primers. Target bp Forward Primer Reverse Primer Coccidian 18$ rDNA 230 188 9F—GTTGGITTCTAGGACTGA 18$ lOR-AGCATGACGTTI‘TCCTATCTCTA Vertebrate 185 rDNA 240 V1 BSF-CTCTTTCGAGGCCCTGTA V18SR-GGGACACTCAGCTAAGAGCA 3. darlingi 1151070 212 HSP7OF-GTGAGGCCGCAAARAAYGA lISP70R-ATGGCGGAGACYTCYTCGG Brilliant® SYBR® Green QPCR Master Mix (Stratagene) was used to quantify the accumulation of double-stranded DNA products in a Stratagene Mx300P Real Time PCR System. Melting curves were performed to assure that the fluorescence was derived from dye intercolating into a specific, homogeneous amplification product. Individual 50p1 reactions contained 25 p1 of SYBR® Master Mix, 19pl of ddH20, lpl of primer mixture comprising lOOnM of each primer, and 5p] of template. Subsequent reactions were performed using identical reagent concentrations in a total volume of 25 pl. To establish that each assay for parasite DNA was quantitative, serial 5-fold dilutions of the purified parasite extracts were amplified. Validation of the assay targeting vertebrate rDNA made use of serial dilutions of crude preparations. Although the absolute concentration of parasite DNA in our reference sample was initially unknown, we deemed quantitative these assays whose linear regression of Ct against relative concentration had an R2 value of >90 and an estimated efficiency between 90 and 110%. Assay reproducibility was determined through replication among extractions and among triplicate qPCR assays of individual extracts. A 10 min pre-incubation at 95°C was followed by 40 cycles of 30 s at 95°C, 1 min at 55°C, and 1 min at 72°C. The subsequent dissociation step included 1 min incubation at 95°C, ramping down to 55°C at a rate of 0.2°C/sec, followed by 81 cycles 55 of incubation where the temperature was increased to 95° by 0.5°C/cycle every 305. Fluorescence was quantified at the end of the annealing, extension, and dissociation steps in each cycle. Each reaction was performed in triplicate to ensure reproducibility. Three wells were used as negative controls; these lacked template but contained all other reagents. The cycle at which fluorescence significantly exceed baseline levels, Ct, was determined using the normalized fluorescence option of the instrument software. To assess the affect of purification on the effective concentration of parasite DNA targets using qPCR, we first assayed extracts containing 1000 merozoites /200pl, as estimated by ocular haemocytometry. By normalizing a second set of pre- and post- purification templates to a lng/ pl, as estimated by UV absorbance spectroscopy, we then assessed how purification affected the relative proportion of total DNA attributable to parasite and host sources. These estimates, when related to the total DNA concentration estimated from UV absorbance spectroscopy, afforded the means to infer absolute concentrations of host and parasite DNA concentrations before and after column purification. Finally, triplicate assays were performed on mixtures of B. darlingi cultures whose purified content comprised 25%, 50% and 75% of the total volume. Statistical analysis. For determining recovery, parasite production, and plaque forming efficiency, analysis of variance (AN OVA) procedure was performed using SAS PC version 8 (SAS Institute, Cary, NC, USA). Statistical significance was assigned to P- values <0.05. Charts were made using the graphical software Si gmaPlot version 8 (RockWare Inc., Golden, CO, USA). 56 RESULTS We purified S. neurona merozoites and B. darlingi tachyzoites by eluting parasite suspensions through PD-lO columns. Such purification was easy to perform and successfully removed visible host cell contaminants from culture media (Figure 2). Figure 2. Photomicrographs of S. neurona merozoites. (A) Crude merozoite preparations before purification. Scale bar = 50 pl. (B) Merozoites (m) completely free of host cell (hc) contaminants and extraneous debris alter purification using PD-10 column. Scale bar = 25 pl. By counting both merozoites/tachyzoites and debris particles, we estimated the mean purity of S. neurona merozoites and B. darlingi tachyzoites at 97.5% and 96.5%, respectively. We also purified the S. neurona_merozoite and B. darlingi tachyzoite suspensions using cellulose filter pads and filter paper. The percentage purity was estimated at 76% and 84% for S. neurona merozoite and 83%, 88% for B. darlingi tachyzoites, respectively. Even though, the amount of host cell contamination observed in the filter paper-purified parasite fraction was relatively small, purification by filter pads and filter papers reduced the yield to 40% and 30% for S. neurona merozoites and 51% and 35%, for B. darlingi tachyzoites. In contrast, ~ 82% of merozoites and 94% of 57 ”VJ tachyzoites of the original suspensions applied to the columns were recovered (Figure 3). Occasionally, merozoite yields were lower when the filter pads or filter papers were washed with PBS instead of EMEM, perhaps because the serum in EMEM retarded or prevented merozoite/tachyzoite attachment to the filter pads or filter papers while washing. We investigated the effect of each purification method on the viability and rate of in vitro development of parasites and by trypan blue dye exclusion assays and by evaluating the rate of parasite production and plaque forming efficiency. Using the trypan blue dye exclusion assay, the estimated viability of S. neurona merozoites and B. darlingi tachyzoites before filtration was not less than 99%; only very minor reductions in viability were observed in column-purified S. neurona (97.6% i 1.1) and column- purified B. darlingi tachyzoites (97.2% i 0.8). In all methods, the purified organisms were found to have variable abilities to penetrate BT cells and to reproduce. However, the parasite production of column-purified parasites was significantly better than in parasites purified by filter pad or filter paper (13.2 vs 7 and 5.5; P=0.003 and 0.001, respectively, df=39 in case of S. neurona and 11.4 vs 5 and 6.2; P=0.007 and 0.008, respectively, df=39 in case of B. darlingi). Additionally, the plaque forming ability of column-purified parasites was significantly greater than that for parasites purified by filter pad or filter paper (38 vs 22.7 and 15.6, P= 0.01 and 0.009; df=39 for S. neurona and 54 vs 19 and 25; P: 0.002 and 0.006, respectively, df=39 for B. darlingi). On the contrary, no significant difference was observed between the crude parasite and column- purified parasites (P < 0.86). 58 - Recovery “/- m Purity % - Merozoite # 0 .00 ..... - Plaquestt m.dddddddddlddddddd0000000000 a Q Q . .....O...°.O...O...C- , .9 O O O b D D O D O tObOOOOOOOOObODOOOOODODQ ’9’ v o v o t o 0' ...n. v o v .9 o ”warn. - . 11111111111 00000000000 120 100 0 0 0 8 6 4 :5 “3 :32 20 Column Pad Paper Purification methods Control 120 % a... e U % # .m new“ w .H a. r c u I... c R P P M _ m - — uuuuuuuuuu"uuuunumnuuuuununu"nun"ununuuuuuuunuuuuuuuuuu EwwmmwwmumwmEmmawwmwfiflafi m .u”aflawHanwfifiwfifiwwfi“Hummmummmmmmmflfl III...-II..-III-IIIIIIIIIIIIIIIIII_ uuuuunuuuaunuuumunnuuuuunuuuuuuuuu_ 0 0 0 0 0 0 0 8 6 4 2 1 Em i 53: r e p a P; .m .m .0 am... Pm m Del. s u CU .IP 0 r. t n o C filter papers compared to unpure (crude) merozoites. Number of parasites is expressed as . All experiments were conducted in triplicate. Figure 3. Purification of Sarcocystis neurona merozoites (A) and Besnoitia darlingi x 106 tachyzoites (B) cultured in Bovine turbinate cells using PD-lO columns, filter pads, and 59 The efficacy of S. neurona merozoite purification using a PD-lO column was also evaluated by SDS-PAGE and western immunoblotting and silver staining assays. Similar but more intense electrophoretic profiles were obtained from proteins extracted from column-purified merozoites as compared to crude S. neurona merozoite preparations (Figure 4). Western blot analyses revealed ~30 and16 KDa antigens in all tested preparations. The immuno-fluorescent staining (IFA) revealed a marked gain in sensitivity and reduction in background in purified parasite preparations. w-hOCD N Figure 4. Silver-stained SDS-PAGE gel of whole-cell protein preparations of Sarcocystis neurona merozoites showing differences in the protein profiles of crude S. neurona merozoites cultured in BT cells and PD-lO column-purified merozoites. M, molecular weight markers (Broad Range Standards; Bio-Rad); Lanes 1 to 3 contain the following amounts of S. neurona protein extract of column— purified merozoites: 1, 2 pl; 2, 5pl; 3, 10pl. Lane 4 contains 15 pl of S. neurona protein extract of un-pure merozoites. The numbers on the lefi indicate the molecular masses of proteins in kilodaltons (KDa) used as standard markers. 60 DNA extracts of column purified B. darlingi merozoites and S. neurona tachyzoites absorbed markedly less light at 260nm than did extracts of crude parasite preparations (Table 2). This absorbance reduction, if caused solely by a reduction in DNA, would indicate a 75% and 90% reduction in the total DNA content of purified B. darlingi and S. neurona extracts. In mixtures of purified and unpurified B. darlingi preparations, the total DNA concentration declined proportionately with the contribution of unpurified merozoites; purified and unpurified extracts from ~250 merozoites yielded an estimated 218 and 790 ng of DNA, respectively (Table 3). These estimated reductions in DNA be exaggerated if crude preparations are disproportionately contaminated with other materials that also absorb light at 260 nm. We preliminarily tested the effect of column purification on the detection of parasite DNA by performing PCRs on extracts obtained from serial dilutions of purified and unpurified S. neurona merozoites and attempting to visualize the resulting product on ethidium bromide-stained agarose gels. PCR primers specific for S. neurona amplified only the DNA isolated from the parasite and a diagnostic band of 334 bp was detected from both purified and un-purified parasite templates. Even though the amount of DNA obtained from 1000 crude S. neurona merozoites was estimated at ~1.2 times that obtained from 1000 pure S. neurona merozoites, solely column-purified extracts supported PCR amplification of a diagnostic marker when such extracts were diluted 10 and 100-fold (Figure 5). 61 Table 2. Total DNA concentration (ng/ pl) in purified and unpurified preparations of DNA estimated to each contains 1000 merozoites/200p] estimated by UV absorbance spectroscopy. Before Purification After Purification Sarcocystis neurona 74.7 3: 17.93 4.5 :i: .50 Besnoitia darlingi 15.83i.12 4.37 i .31 Table 3. Total and parasite DNA content of defined admixtures of crude and column purified preparations of B. darlingi estimated by UV absorbance and qPCR targetting parasite rDNA. Purernpure Observed Expected Observed Expected ng/pl ng/pl* Ct Ct ** 0:4 15.7-15.9 15.8 15.01-15.09 NA 1:3 12.0-12.7 12.94 14.63-14.85 14.86 2:2 9.4-12.3 10.08 14.07-14.47 14.69 3:1 6.0-7.6 7.22 14.00-14.40 14.54 4:0 4.1-4.7 4.36 13.83-14.22 14.40 *Assuming purified and unpurified contain 218 and 790 ng DNA, respectively. "Assuming purified contain 1.57 times as much parasite DNA as crude 62 M12345678 Unpure Pure 1000 100 10 1000 100 10 +ve —ve Figure 5. Analytical sensitivity of PCR analyses using the diagnostic primers JNB25/J D396 on DNA obtained from crude Sarcocystis neurona merozoites cultured in BT cells and PD-10 column-purified S. neurona merozoites. Lanes: M, molecular size marker of a 100bp ladder; 1, 2, 3 unpure merozoites (1000, 100, 10, respectively); 4, 5, 6 column- purified merozoites (1000, 100, 10, respectively); 7, a positive control; 8 a negative control. Numbers on the right and on the left of the gel are DNA fragment sizes (in base pairs). 63 To more precisely quantify the affect of purification on the abundance and purity of parasite DNA extracts, we developed three new quantitative real-time PCR assays. Standard curves prepared from 5-fold serial template dilutions demonstrated an inverse linear relationship between target DNA concentration and the number of amplification cycles necessary to significantly exceed background fluorescence (Ct) (Figure 6A). For assays targeting parasite rDNA and hsp70, this relationship remained linear even when purified extracts were diluted to .00032 of their original concentration. For the assay targeting vertebrate rDNA, the assay remained linear at dilutions as low as .0016 of that present in the original crude extract. Repeated application of these assays resulted in estimated efficiencies of94.6%-110.6%, and R—squared values of 957-999. Thus, each assay supported template quantification over a broad concentration range. We first employed qPCR assays to assess how column purification affected the ability to detect best 188 rDNA. Colmnn purification reduced by more than 130-fold the amount of host rDNA detectable from extracts of 1000 B. darlingi merozoites. A more modest reduction in host DNA, of approximately 2-fold, was observed when crude and column-purified extracts of S. neurona.were compared (Table 4). When total DNA was first normalized to an estimated lng/ pl, purification reduced the amount of host rDNA in B. darlingi preparations by an average of 55-fold (Figure 6B) but did not consistently reduce the amount of host rDNA in S. neurona preparations. Column purification consistently permitted earlier detection of paraSite rDNA amplification products. A 0.79 and 0.65 reduction in Ct was observed for B. darlingi and S. neurona, respectively, representing approximately 1.7 and 1.6 fold increases in the effective concentration of each template. An independent assay targeted to the hsp70 64 gene of B. darlingi resulted in a similar detection improvement subsequent to column purification (Table 5). When the total DNA concentration was first normalized to lng/ul, significantly earlier detection of parasite amplification products occurred in purified samples than in crude samples, irrespective of the locus targeted (Figures 6C-D). The rDNA of purified B._darlingi and S. neurona reached the critical fluorescence in such normalized templates 2.18 and 4.25 cycles earlier, representing approximately 4.5-fold and l9-fold increases in the representation of parasite DNA (Table 6). Finally, qPCR assays were performed on defined admixtures of purified and unpurified B. darlingi templates in order to determine whether incremental improvements in PCR sensitivity resulted when the proportion of purified preparations was increased. Incremental increases in the relative proportion of purified parasites caused parasite DNA targets to be detected earlier (Table 3). 65 Fluorescence (dR) 3 E g 24081012141018202224202330323436380 Figure 6. Amplification plot (fluorescence vs. cycle number) in quantitative PCR assays. (A) Proportional delay in the amplification of parasite rDNA as crude B. darlingi are diluted, in serial 5-fold steps, over 4 orders of magnitude. 66 F Iuorescence (dR ) '8‘ O \' 300.; ..... g ..... ..... g ..... g ..... g ..... g ..... g ..... g ..... g. ..... g mo ..... ..... 1....5...E. ..... ..... ..... ..... ..... ..... ..... 200 ..j ..... ; ..... ..... ..... j ..... j ..... ..... 3. 0 - :1 VMW“! W’ '_; ;_ _ 24081012141618202224262030323430390 Cycles Figure 6 (Cont’d). (B) After normalizing crude (red) and column-purified (blue) extracts to a total DNA concentration of lng/ pl, amplification of host 18S rDNA is delayed. 67 E 5 Fluorescence (dR) Cycles Figure 6 (Cont’d). (C) Note amplification of B. darlingi 18S rDNA is accelerated. 68 F Iuorescenee (GR) 2408101214161820222420283032343638 Cycles Figure 6 (Cont’d). (D) Note amplification of B. darlingi hsp70 is accelerated. 69 Table 4. Effect of column purification on the detection of host rDNA amplification in preparations estimated to contain 1000 merozoites/200p]. Ct Ct ACt Fold Before After difference* Purification Purification Sarcocystis neurona 21.74 :t .387 22.66 i .301 0.92 1.89 Besnoitia darlingi 22.81 i .270 29.88 :t .112 7.07 134.36 Table 5. Effect of column purification on the detection of parasite rDNA amplification in preparations estimated to contain 1000 merozoites/200p]. Ct Ct ACt Fold Before After difference* Purification Purification Sarcocystis neurona 16.15 :t .185 15.36 i: .318 0.79 1.73 Besnoitia darlingi 15.04 i .051 14.39 at .187 0.65 1.57 +Ct —— The critical threshold (Ct) value is the number of cycles required for fluorescence to significantly exceed background. *Assuming PCR product doubles with each cycle in the exponential phase. Table 6. Effect of column purification on the detection of parasite rDNA in lng/ pl total DNA. Ct Ct ACt Fold Before After Purification difference* Purification Sarcocystis neurona 21.42 :t .330 17.17 :t .164 4.25 19.02 Besnoitia darlingi 18.34 :l: .037 16.16 :i: .427 2.18 4.53 70 DISCUSSION Accumulated host cell contaminants can hamper the study of intracellular coccidia. Although purification methods for S. neurona merozoites have been previously described (for example those using syringe filters, 3 pm polycarbonate filters, and calcium ionophores) no highly efficient method suitable for routine processing has been reported for the separation of S. neurona merozoites or B. darlingi tachyzoites fiom tissue culture cells. We are aware of no previous attempts to quantify the affect of such purification methods on the yield and purity of parasite DNA. We have found a simple purification method to be suitable for separating S. neurona merozoites and B. darlingi tachyzoites from cultured and have favorably evaluated the method according to parasite (i) yield and purity, (ii) viability, (iii) SDS- PAGE silver staining and immunoblotting, (iv) immune-fluorescent staining, and (v) quantitative PCR. Our newly developed quantitative PCR assays help evaluate the yield and purity of such parasite preparations. The improved detection of parasite DNA in column- purified preparations may have been caused by the removal of inhibitory substances, the enrichment of parasite-derived DNA as a proportion of the total, or differences in the accuracy of enumerating merozoites in purified and unpurified preparations. Whatever the cause, parasite targets were more readily amplified after column purification. In the case of B. darlingi preparations, contamination by host genomic DNA was markedly reduced as a result of column purification. The feasibility of using a PD-10 column for purification of coccidia from cultured cells with a satisfactory recovery was initially reported by Hemphill et al., (1996) 71 working with N. caninum. In the present study, purification using a PD-10 column proved to be a practical method that provided high yield and highly purified parasite fractions for large—scale purification of parasites. The ability to obtain a pure homogenous population of viable S. neurona merozoites and B. darlingi tachyzoites from large-scale culture facilitates the biological, biochemical, immunological, and genetic studies of these specific stages of these parasites life cycle. This technique appears to have general applicability for purifying coccidia from various cultured cell lines. 72 REFERENCES Dame JB, Greiner E, Ellison SP (2001) In vitro culture and synchronous release of Sarcocystis neurona merozoites from host cells. Vet Parasitol 95:251—261 Dubey JP, Davis SW, Speer CA, Bowman DD, de Lahunta A, Granstrom DE, Topper MJ, Hamir AN, Cummings JF, Suter MM (1991) Sarcocystis neurona n. sp. (protozoa: apicomplexa), the etiologic agent of equine protozoal myeloencephalitis. J Parasitol 77:212—21 8 Dubey JP, Lindsay DS, Kerber CE, Kasai N, Pena HFJ, Gennari SM, Kwok OCH, Shen SK, Rosenthal BM (2001 a) First isolation of Sarcocystis neurona from the South American opossum, Didelphis albiventris, from Brazil. Vet Parasitol 95:295—304 Dubey JP, Lindsay DS, Saville WJA, Reed SM, Granstrom DE, Speer CA (2001b) A review of Sarcocystis neurona and equine protozoal myeloencephalitis (EPM). Vet Parasitol 95 :89—1 31 Dubey JP, Lindsay DS, Rosenthal BM, Sreekumar C, Hill DE, Shen SK, Kwok OCH, Rickard LG, Black SS, Rashmir-Raven A (2002) Establishment of Besnoitia darlingi from opossums (Didelphis virginiana) in experimental intermediate and definitive hosts, propagation in cell culture, and description of ultrastructural and genetic characteristics. Inter J Parasitol 32: 1053—1064 Elsheikha HM, Rosenthal BM, Mansfield LS (2004) Dexarnethasone treatment induces susceptibility of outbred mice to Besnoitia darlingi infection. Parasitol Res (In Press) Ginzinger DG (2002) Gene quantification using real-time quantitative PCR: An emerging technology hits mainstream. Exp Hematol 30:503—512 Hemphill A, Gottstein B, Kaufinann H (1996) Adhesion and invasiOn of bovine endothelial cells by Neospora caninum. Parasitology l 12: 183—197 Klein D (2002) Quantification using real-time PCR technology: applications and limitations. Trends Mol Med 82257—260 Lindsay DS, Dubey JP (2001) Direct agglutination test for the detection of antibodies to Sarcocystis neurona in experimentally infected animals. Vet Parasitol 95:179—186 Mansfield LS, Schott HC, Murphy AJ, Rossano MG, Tanhauser SM, Patterson J 8, Nelson K, Ewart SL, Marteniuk JV, Bowman DD, Kaneene J B (2001) Comparison of Sarcocystis neurona isolates derived from horse neural tissue. Vet Parasitol 95:167—178 Tanhauser SM, Yowell CA, Cutler TJ, Greiner EC, MacKay RJ, Dame J B (1999) Multiple DNA markers differentiate Sarcocystis neurona and Sarcocystisfalcatula. J Parasitol 85:221—228 73 CHAPTER 3 PHYLOGENETIC RELATIONSHIPS OF SARCOC YS T IS NE URONA ISOLATES OF HORSE AND OPOSSUM TO OTHER CYST-FORMING COCCIDIA DEDUCED FROM SSUrRNA GENE SEQUENCES Submitted for publication as original paper: Hany M. Elsheikha 1, David W. Lacher 2, and Linda S. Mansfield " 3 (2004) Phylogenetic relationships of Sarcocystis neurona isolates of horse and opossum to other cyst-forming coccidia deduced from SSUrRNA gene sequences. Parasitology Research (In Review). I Department of Large Animal Clinical Sciences, College of Veterinary Medicine, Michigan State University, East Lansing, MI 48824, USA. 2 Microbial Evolution Laboratory, National Food Safety and Toxicology Center, Michigan State University, East Lansing, MI 48824, USA. 3 Department of Microbiology and Molecular Genetics, College of Veterinary Medicine Michigan State University, East Lansing, MI 48824, USA. 74 ABSTRACT Phylogenetic analyses based on sequences of the nuclear-encoded small subunit rRNA (SSUrRNA) gene were performed to examine the origin, phylogeny, and biogeographic relationships of S. neurona isolates from opossum and horse from the State of Michigan, USA, in relation to other cyst-forming coccidia. A total of 31 taxa representing all recognized subfamilies and genera of Sarcocystidae were included in the analyses. Phylogenies obtained by four tree-building methods were consistent with classical taxonomy. The “isosporid” coccidia Neospora, T oxoplasma, Besnoitia, Isospora lacking stieda bodies, and Hyaloklossia formed a sister group to the Sarcocystis spp. Sarcocystis species were divided into 3 main lineages; S. neurona isolates were located in the second lineage and clustered with S. mucosa, S. dispersa, S. lacertae, S. rodentifelis, S. murz's, and Frenkelia spp. Alignment of S. neurona SSUrRNA sequences of Michigan opossum isolates (MIOP5, MIOP20) with S. neurona sequence of a Michigan horse isolate (MHI8) showed 100% homology. Michigan isolates differed in 30/ 1085 (2.7%) bp from a Kentucky horse isolate (SNS). Additionally, S. neurona isolates from horses and opossums were identical based on ultrastructural features and PCR-RF LP analyses and formed a phylogenetically indistinct group. These findings revealed the concordance between the morphological and molecular data and suggested that S. neurona from opossums and horses generated from the same phylogenetic origin. 75 INTRODUCTION Members of the genus Sarcocystis are among the most common and widespread protozoan parasites of a wide variety of vertebrates worldwide. Studies of this genus may be of economic importance since some of the species are important pathogens of livestock and humans. Sarcocystis neurona is the principal agent of a serious neurological disease in horses in the Americas known as equine protozoal myeloencephalitis (EPM) (Dubey et al., 1991; MacKay, 1997; Dubey et al., 2001). Horses acquire S. neurona by ingestion of the sporocyst stage in feces of opossums in contaminated food or roughage (Fenger et al., 1997). Genetic heterogeneity among S. neurona isolates from different animal hosts has been reported (Mansfield et al., 2001; Rosenthal et al., 2001). This genetic diversity is matched by the wide range of hosts infected with this parasite (Dubey et al., 2001). F urthermore, EPM has a wide range of disease expression in horses. This range may affect the variable response to treatment that is observed in horses infected with S. neurona and suggests that strain variation in the organism may be a factor in the clinical management of EPM. It is likely that S. neurona isolates from different hosts represent a single species based on a number of characteristics, including parasite morphology, histopathology, antigenic reactivity, and most notably the parasitic lifestyle. However, recognition of S. neurona as the single causative agent of EPM in horses and other hosts is possibly an oversimplification. Even though S. neurona displays considerable genetic, clinical, and ecological diversity, this parasite has not been the subject of rigorous phylogenetic treatment of the deeper relationships between isolates in relation to other cyst-forming coccidia. In spite of the dramatic improvement in our understanding of the biology and lifecycle of S. 76 neurona (Dubey et al., 2001), both the phylogenetic position of S. neurona within the genus Sarcocystis and the relationships between S. neurona isolates are ambiguous. The monophyly of Sarcocystis spp. is uncertain partly because of difficulties in assessing the homology of phenotypic characters. Therefore, molecular approaches might provide alternative tools to compare, root, and clarify the phylogeny of these organisms. Studies on the small-subunit ribosomal RNA (SSUrRNA) gene have proved useful in elucidating cryptic diversity among parasites and in addressing phylogenetic questions concerning many protozoa, including Sarcocystis spp. (Hasegawa et al., 1985; Barta et al., 1991; Tenter et al., 1992; Ellis et al., 1994; Slapeta et al., 2003). Most of these studies have questioned the monophyly of the genus Sarcocystis and provided a strong correlation between the phylogenies of Sarcocystis species and those of their definitive hosts. Additionally, the SSUrRNA gene is conserved among species and complete or nearly complete sequences of this gene from many coccidia are available. Consequently, the SSUrRNA gene was selected to examine relationships between S. neurona isolates and other members of the family Sarcocystidae. Although our taxonomic sampling is far from exhaustive, it is representative of the diversity within the Sarcocystidae. We sequenced part of SSUrRNA gene for 3 selected S. neurona isolates from a horse and two opossums from Michigan whose relationships had not been previously examined with molecular data. Sequences from additional taxa representing all subfamilies and genera of Sarcocystidae were included in the analyses to broaden the context in which to interpret SSUrRNA diversity. In this study, we tested the hypothesis of the monophyly of Sarcocystis, with particular focus on S. neurona from the North American opossum (D. virginiana) and 77 horses, and among major Sarcocystidae lineages. Also, the hypothesis that phylogenies of S. neurona and its hosts are consistent with a common history was tested. To investigate these hypotheses we developed and differentiated between axenic cultures of S. neurona isolates obtained from opossum sporocysts and from horse neural tissue using ultrastructure and molecular characterization. Additionally, we explored the practical and evolutionary consequences for molecular-based phylogenetic reconstruction of S. neurona and related coccidia based on SSUrRNA sequence data by a wide range of tree- building methods with reference to evolutionary model strategies and parasite host. Results obtained provided new information regarding the interrelationships within S. neurona isolates and the phylogenetic position of S. neurona isolates from opossums and horses in relation to other coccidia. 78 MATERIALS AND METHODS Taxa sources. A total of 31 operational taxonomic units were included in all analyses. Sequences of 28 taxa, which span all of the currently recognized subfamilies and genera of the family Sarcocystidae, were obtained from GenBank (Table 1). Three taxa were axenic cultures of S. neurona collected in the current study from one horse and 2 opossums from Michigan as follows. One isolate recovered from brain tissue of a naturally infected horse was cultured in equine dermal (ED) cells (ATCC CCL 57, American Type Culture Collection, Rockville, Maryland, USA) as previously described (Mansfield et al., 2001). Two feral opossums were trapped on local farms where horses had died with confirmed EPM. The opossums were humanely euthanatized, necropsied, and the intestines screened for Sarcocystis sporocysts. Sporocysts were recovered, excysted, and cultured in ED cells as described (Murphy and Mansfield, 1999). In all cases, inoculated cultured ED cells were examined every 24 hours for 7 days and weekly thereafter for the presence of Sarcocystis merozoites by microscopy. Once significant cell death was observed, initial cultures were passaged to fresh ED cells, and growth of Sarcocystis was observed during long-term incubation. Morphologic characterization. Microscopic evaluation of Sarcocystis merozoites from ED cultures initially inoculated with infected horse neural tissues and opossum sporocysts were performed using a stained cytospin technique to examine the morphology of the schizonts as described (Elsheikha et al., 2003). Also, merozoites were concentrated via low speed centrifugation (1000 xg), fixed in glutaraldehyde, and routinely processed for electron microscopy (EM) by the MSU Electron Optics Laboratory as described (Elsheikha et al., 2003). 79 Table 1. Details of Apicomplexan taxa included in the study. 80 Classification ' Taxa Accession Sequence Final Intermediate No. Length Host Host Ingroup Sarcocystinae Sarcocystis S. neurona AY518209 b 1085 bp Marsupiala Equidae Sarcocystis S. neurona AY518210 b 1085 bp Marsupiala Equidae Sarcocystis S. neurona AY518211 b 1085 bp Marsupiala Equidae Sarcocystis S. neurona U07812 1803 bp Marsupiala Equidae Sarcocystis S. mucosa AF 109679 1807 bp Marsupiala ? Sarcocystis S. dispersa AF 1201 15 1610 bp Owl Muridae Sarcocystis S. muris . M64244 1805 bp F elidae Muridae Sarcocystis S. rodentzfelis AY0151 11 1593 bp Felidae Muridae Sarcocystis S. Iacerate AY0151 13 1 802 bp Snake Lizard Sarcocystis S. fusiformis U03071 1878 bp Felidae Bovidae Sarcocystis S. buflalonis AF017121 1 884 bp Felidae Bovidae Sarcocystis S. hirsute AF017122 1889 bp Felidae Bovidae Sarcocystis S. hominis AF 006470 1505 bp Primates Bovidae Sarcocystis S. sinensis AF 266959 1405 bp ? Bovidae Sarcocystis S. arieticanis L243 82 1821 bp Canidae Bovidae Sarcocystis S. cruzi AF 01 7120 1818 bp Canidae Bovidae Sarcocystis S. singaporensis AF434054 1801 bp Snake Muridae Frenkelia F. microti AF009244 1631 bp Birds Muridae F renkelia F. glareoli AF 009245 1630 bp Birds Muridae Toxoplasmatinae T oxoplasma T. gondii M97703 1790 bp Felidae Mammals Neospora N. caninum I 271354 1769 bp Canidae Bovidae Besnoitia B. besnoiti AF 109678 1792 bp Felidae Bovidae Besnoitia B. jellisoni AF 291426 1794 bp Felidae Muridae Hammondia H. hammondia AF 096498 1747 bp Felidae Muridae Hyaloklossia H.1ieberkuehni AF 298623 1564 bp Geen frog ? Table 1 (Cont’ d). Cystoisosporinae Cystoisospora I. belli AF 1 06935 1784 bp Human Muridae Cystoisospora I. ohioensis AF 029303 1392 bp Canidae Muridae Cystoz'sospora I. suis ISU97523 1822 bp Pig Muridae Outgroup Eimeridae Eimeria E. tenella U40264 1756 bp Avian ? Cyclospora C. cayetanensis AF111183 1795 bp Human ? Caryospora C. bigenetica AF060976 1751 bp Snake Mammals a Family Sarcocystidae is subdivided into two subfamilies, the Sarcocystinae (2 genera) and the Toxoplasmatinae (5 genera) (Smith and F renkel 2003). b Present study. DNA extraction, amplification, and sequencing. For the three Sarcocystis isolates obtained in this study, merozoites from 3 heavily infected tissue culture flasks were centrifuged and washed with Tris-EDTA buffer. DNA was extracted from these culture-derived merozoites using the Qiagen DNeasy tissue kit. DNA concentration was assessed using a NanoDrop DN-1000 Spectrophotometer and NanoDrop software version 2.5.4. (N anoDrop Technologies, Rockland, DE, USA). Then DNA was tested by S. neurona-specific PCR-RFLP analysis for the presence of the parasite (Tanhauser et al., 1999). Oligonucleotide primers for the SSUrRNA gene were designed from highly Conserved regions detected by alignment of published sequences. The primers are described in Table 2 and were used for both PCR and sequencing reactions. PCR reactions were performed under the following conditions: the total volumes Were 50 p1 containing 1.5 U AmpliTaq DNA Polymerase, 5 pl (10X) reaction buffer, 1.5 mM MgC12, 1 p1 dNTP (1 nM), 1 pl each primer (10 pM), 50 ng template DNA. 81 Amplification was implemented with denaturing at 94°C for 10 min, 35 cycles of denaturing at 92°C for 1 min, annealing at 60°C for 1 min, and extension at 70°C for l min, followed by extension at 72°C for 5 min in a FTC-100 thermocycler (MJ Research Inc.). PCR products were evaluated on 1.5% agarose gels to check for appropriately sized products. Gels were stained with ethidium bromide and visualized by UV transilltunination. Table 2. Names, locations, and DNA sequence of primers used for amplification and sequencing of portion of SSUrRNA gene from Sarcocystis neurona isolates from horses and opossums. Primer name Orientation Position ' Primer sequence (5’-3’) SN—SSUrRNA-Fl Forward 688-708 GAATTCTGGCATCCTCCTGA SN-SSUrRNA -F2 Forward 1182-1201 CTGCGGCTTAATTTGACTC SN-SSUrRNA-Rl Reverse 1285-1305 CATGCACCACCACCCATAGA SN-SSUrRNA -R2 Reverse 1784-1804 CCTACGGAAACCTTGTTACG 3 Numbers indicate positions relative to the S. neurona, SSUrRNA sequence (GenBank accession no. U07812). The amplified products were purified from unincorporated primers and dNTPs with the Qiagen QiAquick PCR Purification Kit and directly sequenced. Some samples were sequenced twice to assure accuracy. Sequencing reactions were carried out on a thermal cycler (FTC-100 thermocycler) with the Big Dye sequencing kit; the programs and reaction were implemented according to the manufacturer’ recommendations, then electrOphoresed in a Beckman CEQ 8000 sequencer (Beckman Coulter Inc., Fullerton, CA, USA). PCR-RFLP assays and SSUrRNA sequencing were also used to distinguish between axenic S. neurona merozoites obtained from horse neural tissues and sporocysts 82 isolated from opossums after long-term growth in tissue cultures. These procedures were repeated six independent times for opossum isolates. Sequence alignment and datablock creation. Sequences acquired from the automated sequencer in the form of electropherograms were aligned, examined thoroughly, and visually checked, and verified. Sequences of the 3 S. neurona isolates obtained in the present study were aligned against 28 previously published SSUrRNA nucleotide sequences from related cyst-forming coccidia obtained from GenBank (see Table 1 for details). The total 31 sequences analyzed represented 7 genera and two subfamilies and were a representative sample of members of the family Sarcocystidae (e. g. species with felids, canids, and marsupials as definitive hosts). Inter-individual variation was taken into account, by including multiple species of some genera and multiple isolates of some species in the analyses. The former were included to resolve relationships within certain genera and to aid in the correct placement of these genera. The latter represented species that were originally sequenced several times due to availability of multiple isolates and to provide confirmation of sequence data. This strategy allows the examination of relative levels of divergence at different taxonomic levels. Sequences were aligned by Clustal X version 1.81 (Thompson et al., 1997), multiple alignment parameters were as follows: gap open penalty 15, gap extension penalty 6.66, transion weight 0.5, delay divergent sequence 30%; finally the alignment was refined by eye. A datablock of the aligned sequences was created in Nexus format. Additionally, the alignment was evaluated in MacClade 4.0b6 (Maddison and Maddison, 2000) with taxon names hidden. Since monophyly can only be assessed on a rooted tree, 83 outgroups were selected based on a higher level phylogeny of the Apicomplexa and according to taxonomic and genetic criteria. Eimeridae represents the most closely related family to the Sarcocystidae (Levine, 1985) and includes parasites such as Genus Eimeria, and the recently identified genera Cyclospora and Caryospora. Therefore, one sequence fi'om each of these 3 genera was chosen and these sequences were used as outgroup taxa in all analyses. Because base compositional heterogeneity is known to affect phylogenetic inference (Lockhart et al., 1994; Yang and Roberts, 1995), base composition and departures from the average base-composition homogeneity across all taxa were evaluated prior to phylogenetic analysis using a )8 analysis. Phylogenetic analyses. Aligned sequences were analyzed using neighbor-joining (NJ), maximum parsimony (MP), maximum likelihood (ML), and Bayesian approaches of phylogenetic inference. All analyses were performed using the simple heuristic search option in PAUP* Version 4.0b10 (PPC) (Phylogenetic Analysis Using Parsimony; Swofford, 2002; Smithsonian Institute) unless otherwise stated. Molecular phylogenies were reconstructed using the NJ method (Saitou and Nei, 1987). NJ tree was constructed for all taxa from the distance matrix for each model tested. The relative distances among different taxa were estimated using uncorrected “p”, F81 (Felsenstein, 1981), F84, Jukes- Cantor (J ukes and Cantor, 1969), Kimura’s two-parameter (K2P) (Kimura, 1980), Hasegawa-Kishino-Yano (HKY85) (Hasegawa et al., 1985), Tajima Nei, Tamura Nei (Tamura and Nei, 1993), and General time reversible (GTR) (Yang, 1994) models. The reliability of the NJ tree was assessed by the bootstrap method with 1,000 replications. We used 95% as the statistically significant value (Efron et al., 1996); however, values greater than 50% were reported since bootstrap values may be conservative estimates of 84 the reliability of groups. In MP analysis, the single gaps and gaps that aligned unambiguously forming indels were treated as “missing data”. Of the total 1993 character, 1646 uninformative characters were excluded. Of the remaining 347 characters, character states were treated as unordered and equally weighted and all were parsimony informative. An initial heuristic search was performed on the aligned dataset (PAUP options including heuristic search, starting tree (start) = stepwise addition, random addition of taxa sequence (10 replicates), branch swapping (swap) = TBR, accelerated transformation [ACCTRAN], and MULTREES on). Bremer support values (Bremer, 1994) were calculated by searching tree space for the shortest tree(s) and decay indices were calculated for each node. Tree length (L), retention index (RI), consistency index (CI), rescaled consistency index (RC), and homoplasy index (HI) were recorded. For evaluation of the robustness of each branch, bootstrap values were calculated using 100 non-parametric pseudoreplicates with heuristic searches employed within each replicate (with the same PAUP options as the initial search and no more than 500 trees saved per replicate). Nucleotide changes were used to determine branch lengths of the tree(s). An additional MP analysis was performed identical to that described above, except that gaps were treated as a fifth base. MP trees(s) were reconstructed as strict consensus and as 50% maj ority-rule consensus tree(s). Character evolution and character state defining clades were analyzed using MacClade Version 4.0b6 (Maddison and Maddison, 2000). NJ and MP trees generated in PAUP were imported to MacClade. All uninforrnative characters were excluded and tree(s) L, CI, and RI were recorded. Character phylogeny was reconstructed and all 85 informative characters were traced to infer the evolutionary histories of the different character states and to see how they are hypothesized to have evolved in the trees. Trees were compared to each other and screened for any differences. Characters that supported the monophyly of the Sarcocystidae and resolved the origin of S. neurona were examined and compared in both trees. For these selected characters, the number of steps, CI and RI were recorded. Analyses under the maximum likelihood criterion were preceded by an evaluation of alternative models of sequence evolution using a fixed tree generated NJ method (Saitou and Nei, 1987) with J ukes—Cantor distances (J ukes and Cantor, 1969). The best- fit model was the one for which additional parameters no longer significantly improved the log-likelihood score, as determined with a likelihood-ratio test (Huelsenbeck and Rannala, 1997). Evaluation of the best-fit model of nucleotide substitution was determined by evaluating the likelihood of 56 alternative nucleotide substitution models using the program Modeltest Version 3.06 (Posada and Crandall, 1998) and PAUP*. The distribution of likelihood scores for all the models expressed as -ln likelihood was obtained (unpublished data). Then, the output score file was imported to Modeltest to compare models by likelihood ratio and Akaike Information Criterion (AIC) tests. A series of likelihood ratio tests were performed for each successive pair of models to select the least complex but most powerful model (Swofford et al., 1996; Sullivan and Swofford, 1997). The significance of increases in likelihood caused by addition of parameters (e.g., allowing for among-site variation in rates) was tested using the hierarchical likelihood ratio test (hLRTs) statistic (—21nA = 2[ln Az—ln in], where M is the likelihood of the 86 restricted model (Felsenstein, 1988). This analysis selected the TrN + I + G model (Tamura and Nei, 1993) with the following parameters (basefreqs = 0.2700, 0.1885, 0.2506, 0.2909; rate matrix = A—Czl, A—G:2.8408, A—Tzl, C—Gzl, C—T:3.9422, G—T:l; proportion of invariable sites (I) = 0.5484, gamma distribution (F) = 0.4988) as the most likely representation of the data. TrN + I + G model is a model with unequal rates for transversions and transitions but with equal rates of all transversions and different rates for each transition. The preliminary ML tree was significantly more likely under the TrN + I + G model than any other model (—1nL= 7745.2168, d.f. = 1, and P < 0.000001) using hLRTs. Akaike Information Criterion test was also used. This test states that a model that minimizes AIC= [—2.log (likelihood)] + [2(number of free parameters)] is considered to be the most appropriate (Akaike, 1974). AIC selected the GTR + I + G as the best-fit model with the following parameters (basefreqs = 0.2661, 0.1933, 0.2542, 0.2865; rate matrix = A-C:0.8235, A—G:3.0090, A—T:1.3590, C—G:0.9l99, C—T:4.1126, G—Tzl; proportion of invariable sites (I)=0.5516, gamma distribution (F) = 0.5076). The preliminary ML tree was significantly more likely under the GTR + I +G model than any other model (—lnL = 7741.5874, AIC = 15503.1748) using AIC test. Subsequent to model evaluation and selection, the ML tree was determined using a heuristic search in PAUP* without topological constraints in which the parameter values under the best-fit model were fixed and a NJ tree was used as a starting point for TBR branch swapping. The resulting tree topology and new parameter estimates were used in a second round of branch swapping to provide the final ML tree; reported parameter values were estimated on this tree. Bootstrap support for nodes in the ML tree was evaluated for 100 pseudoreplicates using TBR branch swapping on starting trees 87 obtained by NJ. Log-likelihood scores for topologies of ML, NJ, and MP trees were calculated using the estimated likelihood parameters and compared using Kishino- Hasegawa test (Kishino and Hasegawa, 1989) in PAUP* with 1000 bootstrap replicates for RELL. To further assess the evolutionary relationships of the coccidia under consideration, Bayesian analysis were performed using MrBayes program, Version 3.0b4 (Huelsenbeck and Ronquist, 2001) with general-time-reversible + gamma + invariants (GTR + G + 1) model of sequence evolution and four heated Markov chain Monte Carlo (MCMC) sampling and a default temperature set to 0.2. Parameters in MrBayes were as follows: nst = 6, rates = gamma, basefreq = estimate. An initial 1000 generations were run to estimate the time per generation. Then, a second run was done with ngn set to a value that would result in an approximately 10 minute run. After the 10 minute run was completed, the likelihood value converged to a stable value. Finally, runs were allowed to proceed for 500,000 generations and trees were sampled every 100 generations. The first 100,000 generations (1000 trees) were discarded as bum-in, and a 50% majority-rule consensus tree was constructed of the remaining sampled trees using PAUP* to obtain posterior probability estimates for each node on the tree. 88 RESULTS Differentiation between opossum and horse isolates. The development of axenic strains of S. neurona from horse and opossums was achieved. S. neurona isolates from horse and opossum were identical based on ultrastructure features and by PCR- RFLP and sequence analyses. S. neurona merozoites were harvested successfully from ED cell cultures inoculated with opossum’ sporozoites. S. neurona ED cultures inoculated with neural tissues of naturally infected horse also produced ample merozoites. Parasites from both hosts underwent multiple cycles of schizogony in ED cells for long-term. EM showed that the ultrastructure characteristics of merozoites resulting from ED culture of opossum sporocysts were identical to axenic S. neurona merozoites obtained from ED culture of horse neural tissues (Figure 1). 1 .9 V5 Figure 1. TEM of S. neurona merozoites cultured in equine dermal (ED) cells and budding out of the residual body (RB). They had conoids (CO), micronemes (MN), and dense granules (DG). Note the presence of dividing nucleus (N) and nucleolous (N O) and the absence of rhoptries, which exist in S. falcatula. Scale bar = 1pm. 89 The S. neurona-specific PCR-RFLP assays produced positive tests on merozoites obtained from opossum’ sporocysts and horse neural tissues cultured in ED cells, but did not amplify the control S. falcatula DNA. Later testing with other S. neurona-specific PCR-RF LP assays even after long-term grth of the organism in cultured ED cells also confirmed identity of the horse and opossum isolates as S. neurona and not S. falcatula (Tanhauser et al., 1999). We tested an aid for distinguishing whether more than one species of Sarcocystis is present in a particular culture, which involved amplification and sequence comparisons of SSUrRNA gene from 6 replicate samples from one opossum isolate. We obtained 100% homology between all 6 replicate sequences from each isolate (Unpublished data). SSUrRNA sequences of S. neurona isolates from opossums and horses were aligned with the previously published sequences of other cyst-forming coccidia(Table1). The SSUrRNA sequences of the two opossum isolates (MIOP5 and MIOP20) and Michigan horse isolate (MIT-18) were identical, and they differed from the sequence of the Kentucky horse isolate (SN5) only in 30 nucleotide positions. Dataset and sequence attributes. The partial sequences of the SSUrRNA gene of three selected S. neurona isolates from a horse and 2 opossums from Michigan were determined (GenBank accession numbers AY518209, AY518210, AY518211). The final SSUrRNA sequence aligrunent used in all analyses consisted of 1,993 bp including gaps and missing data. Of these 466 (23.4%) were invariant, 1646 (82.6%) were parsimony uninformative, and 347 (17.4%) were parsimony informative. All insertion—deletion events were coded as missing data ("7") for purposes of phylogenetic analysis. The final alignment contained sequences from 31 taxa, 28 for the ingroup including the three Michigan isolates and three for the outgroup taxa (Table l). The average base frequencies 90 among all taxa were reasonably uniform, with a slight bias towards A and T (A, 0.230; T, 0.339; C, 0.226; G, 0.203). A chi-square test of homogeneity of base frequencies across taxa showed that none of the taxa had a significant departure from expected base- composition values (for all taxa chi-square (x2) = 200.858; df = 90; P = 0.0005). ()8 value determined using the two-way test of independence implemented in PAUP“). Phylogenetic inferences. Four tree-building methods were used. One of these was a distance-based algorithmic method (NJ) and the other three methods were with optimality criteria, MP, ML, and Bayesian. The PAUP* 4.0b10 computer program (Swofford, 2002) was used for all methods but Bayesian, and employed heuristic searches with global branch-swapping for those methods with an optimality criterion. Representative trees for SSUrRNA gene based on NJ, MP, ML, and Bayesian are shown (Figures 1—3). These trees represent the simplest analyses that resulted in the most resolved trees. NJ phylogenies of SSUrRNA were constructed for the three S. neurona isolates sequenced in this study (MIOP5, MIOP20, and MIT—18) plus 28 published sequences. Phylogenies were rooted using E. tenella, C. cayetanensis, and C. bigenetica as outgroup taxa. NJ phylogenies were constructed using one distance-uncorrected and eight distance-correction models described in Methods. NJ analyses identified 4 major groups and high resolution was seen among all groups: (1) group I included Isospora spp., Besnoitia spp., N. caninum, T. gondii, H. hammondi, and H. lieberkuehm'. Besnoitia was a sister group to a group containing T. gondii, H. hammondi, and N. caninum. Isospora spp. were monophyletic and formed a sister group to the group that encompassed Besnoitia spp., T. gondii, H. hammondi, and N. caninum. Hyaloklossia Iieberkuehni 91 (from European green frog) was the sister to this combined group I. (2) group 11 included F renkelia spp. and non-ruminant Sarcocystis such as, S. neurona, S. mucosa, S. muris, S. rodentifelis, S. dispersa, and S. lacertae and was strongly supported by bootstrap analysis (93% of tress), (3) group 111 included ruminant Sarcocystis spp., and (4) group IV contained S. singaporensis alone, which always formed a separate branch and was a sister species to the carnivore-ruminant Sarcocystis spp. (group III). Taxa in group III were further divided in agreement with their definitive host specificity into two subgroups: (A) Sarcocystis species forming microcysts, with cats as definitive hosts (S. buffalom's, S. hirsuta, S. fusiformis, S. sinensis, and S. hominis) and (B) Sarcocystis species forming macrocysts, with dogs as definitive hosts (S. cruzi and S. arieticanis). In general, all groups were well supported, for example, the Sarcocystis spp. groups 11 and III were strongly supported by parsimony (93% and 100%, respectively) and NJ bootstrap analyses (99% and 95%), respectively. The relationships within group II were not consistent in all models used. The models F 81, Tamura Nei, Tajima Nei, and GTR gave one tree topology. The J ukes-Cantor and uncorrected “P” distance models each gave a different topology. The F84, K2P, and HKY methods gave a similar topology to each other but different from other models. However, those groups that had strong bootstrap support were found consistently using all models. The tree topology obtained with the Tajami Nei model is shown in Figure 1, since more complex models gave substantially similar results. Unweighted MP analysis of 347 phylogenetically informative characters of the SSU rRNA gene with gaps coded as missing data resulted in 117 most parsimonious trees (MPTS), with a length of 809, a CI of 0.648, R1 of 0.812, RC of 0.526, and HI of 0.352. 92 The differences in the topologies of the equally parsimonious trees were primarily associated with uncertainties in the relationships among taxa in group II, which contained S. neurona. A strict consensus of these 117 most parsimonious trees (MPTS) is depicted in Figure 1. The resulting MP trees showed identical topologies to that obtained in the NJ analysis with the Tajami Nei model and hence only strict consensus MP tree is shown. However, bootstrap values for the NJ analyses were shown on the tree (see Fig. 2 legend). When gaps were coded as a fifth character, the MP analysis yielded 3 MPTS (L = 1310). A comparison of the topologies and bootstrap supports of groups obtained after both procedures showed that resolution within all groups did not change but revealed only one difference. When gaps were coded as a fifth character, MP trees showed a different topology for taxa in group II which were further split into 2 distinct subgroups, one subgroup contained 1. ohioensis, H. lieberkuehni, S. rodentifelis, S. dispersa, F. microti, and F. glareoli and the second subgroup contained S. neurona, S. mucosa, S. Iacertae, and S. muris. 93 79 92 98 d1 95 2 Besnoitia besnoiti (AF109678) \ Besnoitia jellisoni (AF 291426) ammondia hammondi (AF 096498) d(Is’bxoplasma gondii (M97703) d2 Neospora caninum (A127l354) dllsospora suis (U97523) Group I Hyaloklossia lieberkuehni (AF 29) J s]: d7 100 8 89 d13 - 10 changes 94 100 f Isospora ohioensis (AF029303) d8 "’0 Isospora belli (AF10693 5) Sarcocystis neurona, MIH8 98 fircocystis neurona, MIOPS 84 Sarcocystis neurona, MIOP20 Sarcocystis neurona (U07812) - Sarcocystis mucosa (AF 109679) — Sarcocystis lacertae (AY0151 13) flfqugrcocystis mum's (M64244) '00 Sarcocystis rodentifelis (AY015) {dfrenkelia microti (AF 009244) Frenkelia glareoli (AF009245) ‘ Sarcocystis dispersa (AF1201 15) d7 100 dStarrcoqystis buflalonis (AF 0 l 7 120) D :33 0° Sarcocystis hirsuta (AF017122) Group II Sarcocystis fusiformis (U 03071) d 6Sarcocystis sinensis (AF266959) 13: ("30° Sarcocystis hominis (AF006470) Sarcocystis cruzi (AF017120) Sarcocystis arieticam's (L243 82) J Sarcocystis singaporensis (AF43) Group IV 9 d4 Cyclospora cayetanensis (AFl l 1 l) Eimeria tenella (U40264) Caryospora bigenetica (AF 060976) Outgroup Group 111 Figure 2. Strict consensus of 117 most parsimonious trees inferred from analysis of 347 phylogenetically informative characters of the nuclear SSUrRNA gene of 17 Sarcocystis spp. and 11 related taxa and rooted with three Outgroup taxa. Neighbor-joining (NJ) tree topology based on Tajami Nei model was almost consistent with the MP trees. Values above and below the nodes represent bootstrap resampling results (% of 100 replications for MP analysis) and (% of 1000 replications for NJ algorithm), respectively. Numbers at the nodes (d) indicate decay indices. Bootstrap values are reported only for clades present in >50% of replicates. GenBank accession numbers of the SSUrRNA sequences of the organisms are given in parentheses. To the right taxonomic affiliation is presented. Branch lengths are proportional to distance. 50% majority role tree has identical topology. 94 ML analyses were employed to allow a more complex model of nucleotide evolution, including substitution rates and probabilities of nucleotide change, to be developed and used to infer the most likely phylogenetic relationships among taxa. The best-fitting models for the SSUrRNA sequences recovered by hLRT and AIC analyses were found to be TrN and GTR, respectively, plus parameters to correct for invariant and rate variation among sites. Identical trees of likelihood (—lnL) 7735.24808 and 7884.84486 were obtained by TrN + I + G and GTR + I + G, respectively. The resulting tree from GTR +1 + G analysis was similar to that found in the strict consensus MP and NJ analyses, with minimal topological differences in taxa in group II (Figure 3), but the tree that resulted from TrN + I + G analysis was different from the trees that resulted from NJ, MP and ML analysis using GTR + I + G in the relationships between taxa in the second group. In the ML tree with TrN + I + G model, the taxa in group II were split into two subgroups; one subgroup contained S. dispersa, S. mucosa, F renkeli microti, and F. glareoli, while the second subgroup contained S. neurona, S. Iacertae, S. muris, and S. rodentifelis. Log-likelihood scores of the trees were calculated using the estimated likelihood (hLRT and AIC) parameters and compared by an Kishino-Hasegawa test (Kishino and Hasegawa, 1989) as implemented in PAUP“ in order to test whether the maximum likelihood tree topology was any better or worse than the other tree topologies found in NJ and MP analyses. The likelihood scores for all trees with each best-fit model used were obtained (Table 3). The Kishino-Hasegawa test indicated that none of these trees are statistically significantly different from each other (P>0.05). 95 100 Besnoitia besnoiti (AF109678) Besnoitiajellisoni (AF291426) Hammondia hammondi (AF 096498) Toxoplasma gondii (M97703 0) Neospora caninum (AJ271354) 100 100 Isospora suis (U97523) — . Isospora ohioensis (AF029303) Isospora belli (AF 1 06935) Hyaloklossia lieberkuehm' (AF29) Sarcocystis neurona, MIH8 100 100 Sarcocystis neurona, MIOPS Sarcocystis neurona, MIOP20 Sarcocystis neurona (U07812) Sarcocystis mucosa (AF 109679) 00 F renkelia microti (A F009244) Frenkelia glareoli (AF 009245) Sarcocystis dispersa (AF 1201 15) Sarcocystis lacertae (AY01 5 l 13) Sarcocystis muris (M64244) Sarcocystis rodentifelis (AY015) 100 Sarcocystis buflalonis (AF01712) ‘— , 10° Sarcocystis hirsuta (AF017122) Sarcocystisfusiformis (U03071) Sarcocystis sinensis (AF266959) Sarcocystis hominis (AF006470) Sarcocystis cruzi (AF 0 l 7 120) Sarcocystis arieticanis (L243 82) 100 100 Sarcocystis singaporensis (AF43) 33 E: Cyclospora cayetanensis (AF 1 l l 1) 10° Eimeria tenella (U40264) L— C aryospora bigenetica (AF060976) — 0.01 substitutions/site Figure 3. Maximum likelihood phylogeny of SSUrRNA among the Sarcocystidae, rooted with three outgroup taxa constructed with a GTR + I + G model of substitution, (- lnL: 7885) heuristic search, random stepwise addition, TBR, equal character weight, with MULPARS in effect. The log likelihood score of the phylogeny is -7884.844. Numbers above nodes represent percent bootstrap support of 100 non-parametric replications. Branch lengths are proportional to the hypothesized amount of inferred evolutionary change (ML distances), as shown by the scale bar. Sequences of the isolates in bold are obtained in this study. Sequences of the isolates in bold are obtained in this study. 96 Table 3. The log likelihood scores (— 1n likelihood) of the topologies of ML, NJ, and MP trees of the SSUrRNA under best-fit models of nucleotide evolution selected by LRT and AIC testing using program modeltest. TrN = (Tamura and Nei, 1993), GTR = (Yang, 1994). Likelihood test Model of nucleotide Tree score evolution ML MP NJ Likelihood ratio TrN + I + G 7735.248 7736.325 7745.216 (LT) Akaike Information GTR + I +G 7884.844 7885.570 7895.448 Criterion (AIC) The pairwise number of nucleotide substitution differences detected among the aligned SSUrRNA sequences from the 31 species of coccidia are shown in Table 3. There were fewer nucleotide differences between the SSUrRNA genes of S. neurona and S. mucosa than there were between S. mucosa and other Sarcocystis spp. analyzed. There were few nucleotide differences between the sequences of T. gondii, N. caninum, and H. hammondi thus explaining the polytomy observed between them. 97 :0 2.0 000 :0 00.0 3.0.0 00.0 3.0 00.0 who :0 «0.0 00.0 3.0 00.0 3.0 3.0 v0.0 30 00.0 00.0 000 $0 and N00 no.0 00.0 W00 3.0 00.0 8.3350300 . _ m - 30 30.0 330 330 33.0 03.0 33.0 03.0 030 030 .00 30 30.0 300 300 00.0 30.0 30.0 000 000 00.0 30.0 30.0 00.0 00.0 000 00.0 30.0 300 0:05.: .m .03 - 3.0 330 30 030 03.0 33.0 33.0 33.0 33.0 30.0 00.0 300 300 00.0 000 000 30.0 030 030 03.0 30.0 300 300 000 :0 30.0 300 30.0 3330008030... 0 .03 - 33.0 30.0 30 3:0 3:0 300 30.0 30 33.0 30 330 330 330 33.0 33.0 03.0 03.0 03.0 03.0 20 20 0.0 :0 :0 :0 3.0 20 3003000.: .03 .33 - 33.0 33.0 _30 .30 .30 3.0 33.0 33.0 03.0 03.0 33.0 33.0 330 33.0 030 3.0 3.0 3.0 33.0 030 33.0 03.0 03.0 03.0 03.0 03.0 35.800003 .0103 - :0 33.0 03.0 33.0 03.0 03.0 3.0 33.0 3.0 33.0 33.0 03.0 33.0 33.0 33.0 330 33.0 33.0 03.0 33.0 000 0:0 0:0 000 03.0 332820.20 .0. .03 - 33.0 :0 33.0 03.0 330 33.0 3.0 33.0 3.0 33.0 3.0 33.0 :30 33.0 33.0 33.0 03.0 :0 03.0 :0 000 :0 00.0 03.0 RE... .0. .33 - 00.0 33.0 03.0 33.0 33.0 33.0 3.0 33.0 33.0 330 3.0 33.0 030 03.0 03.0 30 30 30 :0 30 3.0 :0 00.0 3.3.0500 .3 .03 - 33.0 33.0 3.0 3.0 33.0 3.0 330 03.0 03.0 3.0 3.0 3.0 3.0 3.0 30 $0 30 0.0 :0 :0 :0 03.0 035530133 - 00.0 00.0 3.0 3.0 33.0 33.0 33.0 33.0 03.0 03.0 .30 _3.0 _3.0 03.0 3:0 000 03.0 330 330 030 30 0300335013333 - _0.0 33.0 03.0 33.0 33.0 33.0 33.0 33.0 33.0 _30 .30 .30 :0 000 30.0 00 0:0 000 03.0 03.0 25.3 .0. :3 - 030 030 30 330 33.0 33.0 _30 _3.0 .30 .30 _30 000 30.0 :0 03.0 00.0 00 33.0 03.0 3.003%..00103 - 30.0 30.0 :0 0.0 2.0 30.0 00 30.0 30.0 30.0 03.0 _30 03.0 00 3.0 0.0 :0 3.0 02233.00 .0. .0_ - 30.0 :0 :0 :0 30.0 30.0 30.0 30.0 30.0 03.0 _30 03.0 3.0 :0 :0 :0 :0 08330. K .3. - :0 0:0 3:0 00 30.0 00.0 00.0 00.0 03.0 33.0 03.0 :0 3.0 :0 :0 3.0 320.0: m .: - 30.0 20 3:0 20 30.0 30.0 30.0 33.0 30 3.0 33.0 33.0 33.0 33.0 33.0 30020302 .0. 0_ - :0 20 0:0 :0 :0 :0 33.0 30 3.0 33.0 33.0 33.0 :30 30 3.250.012 - 0.0 2.0 :0 :0 :0 3.0 03.0 33.0 33.0 33.0 33.0 33.0 33.0 00:82 .m .3 - 00.0 30 :00 000 03.0 .30 _30 :0 :0 :0 :0 :0 3.805 .0. .3_ - 00.0 00.0 00.0 _30 33.0 33.0 03.0 03.0 03.0 00 0.0 33va 0:008: .013. - 00.0 00.0 330 03.0 330 33.0 3.0 3.0 330 33.0 830920000000 .01: - 00.0 33.0 03.0 33.0 33.0 30 30 33.0 33.0 300:): 0000.0: .3 .0. - 330 03.0 33.0 33.0 30 30 33.0 33.0 33:50 Essa. .0. .0 - .00 30.0 00.0 0.0 00.0 00 2.0 .500 .3 .3 - 00.0 3.0 :0 :0 :0 30 3.200030 3.0 - :0 :0 0.0 0.0 :0 333.30 - 00.0 00.0 300 30.0 5:508 .2 .3 - 00.0 000 30.0 30203 g .3. - 30.0 30.0 00090500 .33 .3 - 30.0 3:03.503 .33 .3 i QCNOEMNQ .m .— ..3 03 .03 .33 .33 .03 .33 .03 .33 .33 ..3 .03 .0_ .2 .: .0. .3. .3 .3. .3. .: .0. .0 .3 .3 .0 .3 .v .3 .3 ._ 532.380 05.5%? 00 3033053 00:80.6 0.30:0 08088035 .3 00583820 363:8wco .5300 0:0 .003 0.2038950. 00 38:30:33 90%). Few topological differences were found in MP analysis with gaps coded as a fifih base, in ML analysis using the TrN + I + G model; in the Bayesian method only in the positions and relationships of the taxa in group II. In the present study, we examined the relationship between S. neurona isolates and other Sarcocystis spp. and the question of whether Sarcocystis species having non- ruminants as intermediate hosts form a monophyletic group. The previous phylogenetic analyses of one S. neurona isolate (SN5) from a horse in Kentucky based on SSUrRNA gene sequences indicated close but incompletely resolved phylogenetic position of S. neurona (Fenger et al., 1994). These authors found that S. neurona isolate SN5 clustered with T oxoplasma gondii. In this study, S. neurona isolates from opossums and horses occupied different positions in the SSUrRN A phylogenetic tree rather than previously reported. Previous studies questioned the monophyly of the genus Sarcocystis based on morphological criteria alone. The species of this genus clustered into three distinct sister groups, one of which (group II) shared a common ancestor with F renkelia spp. The close relationship of S. neurona, S. mucosa, S. muris, S. dispersa, S. Iacertae, and S. rodentifelis with two F renkelia spp. supported the conclusion that Sarcocystis is paraphyletic, in agreement with results obtained by Jenkins et al.,(l999) and Mugridge et al., (1999). There are biological and epidemiological differences between Sarcocystis spp. and F renkelia spp. and their designation as Sarcocystis, which encyst in muscles, or 103 F renkelia, which encyst in the central nervous system, suggest retaining the practically useful generic distinctions. To date, the numbers of experimental and phylogenetic studies of members of both genera are limited, therefore, deferring taxonomic changes as suggested by Smith and Frenkel, (2003). Phylogenetic analysis provides one of numerous frameworks for the interpretation of biological data. Group II of the Sarcocystidae, which contains S. neurona, consistently included species with a diverse range of hosts. This might explain partly the inconsistency in the positions of taxa in this group. The complete life cycle of S. neurona is almost known with opossums as the definitive host and horses and other vertebrates as aberrant intermediate hosts (Dubey et al., 2001). Other Sarcocystis spp. that clustered closely with S. neurona are known to cycle between cats, marsupials, lizards, birds, and mice. In all analyses, S. neurona isolates clustered closely with S. mucosa, S. muris, S. dispersa, S. Iacertae, and S. rodentifelis, forming a monophyletic group. Similar relationships between these Sarcocystis spp. based on SSUrDNA alignment were also noted by Jenkins et al., (1999). They mentioned that the group that contained S. neurona, S. mucosa, S. muris, F renkelia, and Sarcocystis sp. (from a rattlesnake) was less robust and the position of S. muris and Sarcocystis sp., in particular, was unstable. A more recent study by Slapeta et al., (2003) gave also similar results. The weight of evidence from parasite biology indicates that the grouping of S. neurona isolates with S. mucosa, S. muris, S. dispersa, S. Iacertae, and S. rodentzfelis described here (i.e. a monophyletic group radiating from a common ancestor) might be a reasonable conclusion. Incongruence in co-evolution, as shown by some lizard parasites, may reflect the existence of a common ancestor with low host specificity. However, more data from 104 representative taxa are needed, especially from Sarcocystis spp. that infect reptiles and marsupials to resolve the relationships among the branches in this region of the tree. The lack of resolution in that part of the tree containing F renkelia, S. neurona, S. mucosa, S. muris, S. dispersa, S. Iacertae, and S. rodentz'felis might also indicate slow rates of evolution. The fourth group contained only S. singaporensis, which has a snake-rodent life cycle (Dolezel et al., 1999; Jenkins et al., 1999). S. singaporensis always formed a separate branch and arose as a sister species to the carnivore-ruminant Sarcocystis spp. in all analyses. A similar observation was made by Slapeta et al., (2002). Trees obtained by NJ and MP analyses were topologically identical. However, for a complete comparison the changes and behavior of some characters in both trees were also compared. CI, RI, and RC are statistics describing how a particular character behaves within a particular tree. Based on the values of CI, RI, and RC in both trees (= 1 in all cases), the evolution of character 297 in the NJ and MP trees was completely consistent and gave us a preliminary clue about the origin of S. neurona. Additionally, both sets of characters in each tree sharing many homologous characters with very few different characters. Character 297 and the similar characters between the two trees produced trees that have a reasonable biological meaning and reflected the evolutionary relationships of S. neurona and support the hypotheses that S. neurona evolved as a subset fi'om other Sarcocystis spp., and the monophyly of Toxoplasmatinae and Sarcocystinae. The Modeltest program gave two different results under two different criteria, hLRTs and AIC. Two models of nucleotide substitutions were used, in order to test the sensitivity of the analysis to the source of variability. As expected, the most complex 105 model GTR + I + G obtained by the AIC test was better than the TrN + I + G model obtained by LRT. Even so, the ML tree generated by the TrN + I + G model had a higher likelihood score (= lowest -In L) than the ML tree generated by the GTR + I + G; the latter tree was completely consistent topologically with trees obtained by NJ and MP with gaps coded as missing; furthermore, it agreed with the conventional taxonomy, parasite biology, and previous studies. Trees obtained by the ML plus TrN + I + G model, the MP tree with gaps treated as a fifth base, and the Bayesian tree disagreed with all other methods in the arrangement and relationships between taxa in group 11. Many studies have favored co-evolution of Sarcocystis spp. with the final rather than the intermediate host (Doleiel et al., 1999; Slapeta et al., 2001; Slapeta et al., 2002). Our results are consistent with this scenario. The exact geographic origin of S. neurona is unknown, but it may not be indigenous to the New World. The similarity of S. neurona and S. mucosa suggests that S. neurona and S. mucosa have a common evolutionary history. The genetic distance between S. neurona and S. mucosa was found to be 4.3% (Table 3). Escalante and Ayala, (1995) used a general evolutionary rate of 0.85% sequence divergence/ 100 millions of years (Myr) for divergence of the Apicomplexan parasites in general for the entire SSUrDNA and 2-4% sequence divergence/ 100 Myr for the hypervariable regions. These findings indicate that these Sarcocystis spp. diverged from each other at least 100 Myr ago. A slower rate of nucleotide-sequence divergence in marsupials of 0.31% per 100 Myr was proposed by Westennan et al., (1990), which was based on an earlier estimate of 103-128 Myr before present for the separation of Didelphid (American opossum) and Australian marsupials (Richardsons, 1988). The fact the divergence time of S. neurona and S. mucosa and the divergence time of their hosts 106 are similar is consistent with the hypothesis that these species co-evolved with their definitive hosts, carnivorous marsupials (J akes, 1998; Rosenthal et al., 2001). Finally, it is thought that the continuous land connections between South America and Australia started to separate 140 Myr ago (Veevers, 1991). Recently, S. neurona has been shown to be transmitted from South to North America by opossums through the Panamanian isthmus (Rosenthal et al. 2001). Therefore, it is possible that S. neurona may have been derived from S. mucosa as suggested by J akes (1998). The geographic separation between Americas and Australian continents might have precluded gene flow between S. neurona and S. mucosa, which might cause alleles to be distributed unevenly among organisms. This isolation might have followed by founder effect and genetic drift that enabled S. neurona to be genetically distinct from S. mucosa. Further phylogenetic analyses of the SSUrRNA and additional polymorphic genes from a large number of these two organisms and other Sarcocystis from the three continents are needed to address these issues more deeply. 107 REFERENCES Akaike H (1974) Information theory and an extension of the maximum likelihood principle. In Proceedings of the 2nd International Symposium on Information Theory (ed. Petrov, B.N. & Csaki, F.), pp. 267—281. Akademia Kiado, Budapest, Hungary Barta JR, Jenkins MC, Danforth HD (1991) Evolutionary relationships of avian Eimeria species among other apicomplexan protozoa: monophyly of the Apicomplexa is supported. Mol Biol Evo 82345—355 Barta JR, Martin DS, Carreno RA, Siddall ME, Profous-Juchelkat H, Hozza M, Powles MA, Sundermann C (2001) Molecular phylogeny of the other tissue coccidia: Lankesterella and Caryospora. J Parasitol 87:121—127 Box ED, Meier J L, Smith JH (1984) Description of Sarcocystisfalcatula Stiles, 1893, a parasite of birds and opossums. J Protozool 31 :521—524 Bremer K (1994) Branch support and tree stability. Cladistics 10:295—304 Carreno RA, Barta JR (1999) An eimeriid origin of isosporid coccidian with stieda bodies as shown by phylogenetic analysis of small subunit ribosomal RNA gene sequences. J Parasitol 85:77—83 ' Carreno RA, Schnitzler BE, J effries AC, Tenter AM, Johnson AM, Barta JR (1998) Phylogenetic analysis of coccidia based on 188 rDNA sequence comparison indicates that Isospora is most closely related to T oxoplasma and Neospora. J Eukaryot Microbiol 45: 184—1 88 Dolezel D, Koudela B, J irku M, Hypsa V, Obomik M, Votypka J, Modry D, Slapeta JR, Lukes J (1999) Phylogenetic analysis of Sarcocystis spp. of mammals and reptiles supports the coevolution of Sarcocystis spp. with their final hosts. Inter J Parasitol 29:795—798 Dubey JP, Dabis SW, Speer CA, Bowman DD, de Lahunta A, Granstrom DE, Topper MJ, Hamir AN, Cummings J F, Suter MM (1991) Sarcocystis neurona n. sp. (Protozoa: Apicomplexa), the etiologic agent of equine protozoal myeloencephalitis. J Parasitol 77:212—218 Dubey JP, Lindsay DS (1998) Isolation in immuodeficient mice of Sarcocystis neurona from opossum (Didelphis virginiana) faeces and its differentiation from Sarcocystis falcatula. J Parasitol 29:1823-1828 Dubey JP, Lindsay DS (1999) Sarcocystis speeri n. sp. (Protozoa: Sarcocystidae) from the opossum (Didelphis virginiana). J Parasitol 852903—909 108 Dubey JP, Lindsay DS, Saville WJ, Reed SM, Granstrom DE, Speer CA (2001) A review of Sarcocystis neurona and equine protozoal myeloencephalitis (EPM). Vet Parasitol 95 :89—1 3 1 Dubey JP, Speer CA, Lindsay DS (1998) Isolation of a third species of Sarcocystis in immunodeficient mice fed feces from opossums (Didelphis virginiana) and its differentiation from Sarcocystisfalcatula and Sarcocystis neurona. J Parasitol 84:1158— 1 164 Efron B, Halloran E, Holmes S (1996) Bootstrap confidence levels for phylogenetic trees. Proc Natl Acad Sci U S A. 93:13429—13434 Ellis JT, Holmdahl OJ, Ryce C, Njenga J M, Harper PA, Morrison DA (2000) Molecular phylogeny of Besnoitia and the genetic relationships among Besnoitia of cattle, wildebeest and goats. Protist 151 :329—336 Elsheikha HM, Saeed MA, Fitzgerald SD, Murphy AJ, Mansfield LS (2003) Effects of temperature and host cell type on the in vitro growth and development of Sarcocystis falcatula. Parasitol Res 91 :22—26 Escalante AA, Ayala FJ (1995) Evolutionary origin of Plasmodium and other Apicomplexa based on rRNA genes. Proc Natl Acad Sci U S A 92:5793-5797 F enger CK, Granstrom DE, Langemeier J L, Gajadhar A, Cothran G, Tramontin RR, Stamper S, Dubey JP (1994) Phylogenetic relationship of Sarcocystis neurona to other members of the family Sarcocystidae based on small subunit ribosomal RNA gene sequence. J Parasitol 80:966-975 Fenger CK, Granstrom DE, Gajadhar AA, Williams NM, McCrillis SA, Stamper S, Langemeier J L, Dubey JP (1997) Experimental induction of equine protozoal myeloencephalitis in horses using Sarcocystis sp. Sporocysts from opossums. Vet Parasitol 68: 199—2 1 3 Felsenstein J (1981) Evolutionary trees from DNA sequences: a maximum likelihood approach. J Mol Evo 172368—376 Felsenstein J (1988) Phylogenies from molecular sequences: Inference and reliability. Annu Rev Genet 22:521—565 Frenkel JK (1977) Besnoitia wallacei of cats and rodents: with a reclassification of other cyst-forming isosporoid coccidia. J Parasitol 63:611—628 F renkel JK, Dubey JP (1972) Rodents as vectors for feline coccidia, Isosporafelis and Isospora rivolta. J Infect Dis 125:69—72 109 Hasegawa M, Iida Y, Yano T, Takaiwa F, Iwabuchi M (1985) Phylogenetic relationships among eukaryotic kingdoms inferred from ribosomal RNA sequences. J Mol Evo 22:32— 38 Hasegawa M, Kishino H, Yano T (1985) Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J Mol Evo 21 :160—174 Huelsenbeck JP, Rannala B (1997) Phylogenetic methods come of age: testing hypotheses in an evolutionary context. Science 276:227—232 Huelsenbeck JP, Ronquist F (2001) MrBayes: Bayesian inference of phylogeny. Biometrics 17754—17755. Kimura M (1980) A simple method for estimating evolutionary rate of base substitutions through comparative studies of nucleotide sequences. J Mol Evo 16:111—120 Kishino H, Hasegawa M (1989) Evaluation of the maximum likelihood estimate of the evolutionary tree topologies from DNA sequence data, and the branching order in Hominoidea. J Mol Evo 29: 1 70—179 J akes KA (1998) Sarcocystis mucosa in Bennetts wallabies and pademelons from Tasmania. J Wildl Dis 34:594—599 Jenkins MC, Ellis JT, Liddell S, Ryce C, Munday BL, Morrison DA, Dubey JP (1999) The relationship of Hammondia hammondi and Sarcocystis mucosa to other heteroxenous cyst-forming coccidia as inferred by phylogenetic analysis of the 18S SSU ribosomal DNA sequence. Parasitology 119:135—142 J ukes TH, Cantor CR (1969) Evolution of protein molecules. In: H.N. Munro, Editor, Mammalian Protein Metabolism vol. 3, Academic Press, New York, pp. 21—132 Levine ND (1985) Phylum II. Apicomplexa Levine 1970. In: An Illustrated Guide to the Protozoa (Edited by Lee, J .J ., Hutner, S.H. & Bovee, E.C.), pp. 322—374. Society of Protozoologists, Lawrence Lindsay DS, Dubey JP, Horton KM, Bowman DD (1999) Development of Sarcocystis falcatula in cell cultures demonstrates that it is different from Sarcocystis neurona. Parasitology 1 1 8: 227—233 Lockhart PJ, Steel MA, Hendy MD, Penny D (1994) Recovering evolutionary trees under a more realistic model of sequence evolution. Mol Biol Evo 12:605—612 MacKay RJ (1997) Equine protozoal myeloencephalitis. Vet Clin North Am Equine Pract 13:79—96 Maddison WP, Maddison DR (2000) MacClade: Analysis of Phylogeny and Character Evolution, version 4.0. Sunderland, Massachusetts: Sinauer Associates 110 Mansfield LS, Schott HC 2nd, Murphy AJ, Rossano MG, Tanhauser SM, Patterson J S, Nelson K, Ewart SL, Marteniuk JV, Bowman DD, Kaneene JB (2001) Comparison of Sarcocystis neurona isolates derived from horse neural tissue. Vet Parasitol 95:167—178 Mugridge NB, Monison DA, Johnson AM, Luton K, Dubey JP, Votypka J, Tenter AM (1999) Phylogenetic relationships of the genus Frenkelia: a review of its history and new knowledge gained from comparison of large subunit ribosomal ribonucleic acid gene sequences. Inter J Parasit0129z957-972 Murphy AJ, Mansfield LS (1999) Simplified technique for isolation, excystation, and culture of Sarcocystis spp. from opossums. J Parasitol 85:97 9—981 Posada D, Crandall KA (1998) MODELTEST: testing the model of DNA substitution. Bioinforrnatics 14:817—81 8 Richardson BJ (1988) A new view of the relationships of Australian and American marsupials. Austr Mammal 11:71—73 Rosenthal BM, Lindsay DS, Dubey JP (2001) Relationships among Sarcocystis species transmitted by New World opossums (Didelphis spp.). Vet Parasitol 95:133—142 Slapeta JR, Modry D, Votypka J, J irku M, Koudela B, Lukes J (2001) Multiple origin of the dihomoxenous life cycle in sarcosporidia. Inter J Parasitol 31 :413—417 Slapeta JR, Kyselova 1, Richardson AO, Modry D, Lukes J (2002) Phylogeny and sequence variability of the Sarcocystis singaporensis Zaman and Colley, (1975) 1976 serNA. Parasitol Res 88:8 10—8 15 Slapeta JR, Modry D, Votypka J, J irku M, Lukes J, Koudela B (2003) Evolutionary relationships among cyst-forming coccidia Sarcocystis spp. (Alveolata: Apicomplexa: Coccidea) in endemic African tree vipers and perspective for evolution of heteroxenous life cycle. Mol Phylogenet Evo 272464—475 Smith DD (1981) The Sarcocystidae: Sarcocystis, F renkelia, T oxoplasma, Besnoitia, Hammondia, and Cystoisospora. J Protozool 28 262—266 Smith DD, Frenkel JK (2003) Determination of the genera of cyst-forming coccidia. Parasitol Res 91 :384—389 Saitou N, Nei M (1987) The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evo 41406—425 Swofford DL, Olsen GL, Waddell PJ, Hillis DM (1996) Phylogenetic inference. Molecular Systematics, 2nd edn (ed. By Hillis, D.M., Moritz, C. & Mable, B.K.), pp. 407—514. Sinauer Associates, Sunderland, Massachusetts 111 Swofford DL (2002) PAUP“. Phylogenetic Analysis Using Parsimony (* and Other Methods). Version 4. Sinauer Associates, Sunderland, Massachusetts Tamura K, Nei M (1993) Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Mol Biol Evo 1 0:5 12-526 Tanhauser SM, Yowell CA, Cutler TJ, Greiner EC, MacKay RJ, Dame JB (1999) Multiple DNA markers differentiate Sarcocystis neurona and Sarcocystisfalcatula. J Parasitol 85:221—228 Tenter AM, Baverstock PR, Johnson AM (1992) Phylogenetic relationships of Sarcocystis species from sheep, goats, cattle and mice based on ribosomal RNA sequences. Inter J Parasitol 22:503—5 13 Thompson JD, Gibson TJ, Plewniak F, J eanmougin F, Higgins DG (1997) The ClustalX windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 25:4876—4882 Veevers JJ (1991) Phanerozoic Australia in the changing configuration of Proto-Pangea through Gondwanaland and Pangea to the present dispersed continents. Austr Sys Botany 4: 1—1 1 Westennan M, J anczewski DN, O’Brien SJ (1990) DNA-DNA hybridisation studies and marsupial phylogeny. In ‘Mammals from Pouches and Eggs: Genetics, Breeding and Evolution of Marsupials and Monotremes’. (Eds Marshall Graves, J .A., Hope, RM. and Cooper, D.W.) pp. 173—181 Yang Z (1994) Maximum likelihood phylogenetic estimation from DNA sequences with variable rates over sites: Approximate methods. J Mol Evo 39: 1 05—1 11 Yang Z, Roberts D (1995) On the use of nucleic acid sequences to infer early branchings in the tree of life. Mol Biol Evo 12:451—458 112 CHAPTER 4 SARCOCYSTIS NE URONA MAJOR SURFACE ANTIGEN GENE 1 (SAGI) SHOWS EVIDENCE OF HAVING EVOLVED UNDER POSITIVE SELECTION PRESSURE Published as original paper: Hany M. Elsheikha 1 and Linda S. Mansfield 1' 2 (2004) Sarcocystis neurona major surface antigen gene 1 (SA G1) shows evidence of having evolved under positive selection pressure. Parasitology Research 94: 452—459. I Department of Large Animal Clinical Sciences, College of Veterinary Medicine, Michigan State University, East Lansing, MI 48824, USA. 2 Department of Microbiology and Molecular Genetics, College of Veterinary Medicine, Michigan State University, East Lansing, MI 48824, USA. 113 ABSTRACT The major surface antigen gene 1 (SA G1) is conserved among members of Sarcocystidae and may play an important role in parasite pathogenesis. Additionally, generation and selection of different antigenic variants of SA GI has the potential for inclusion in a subunit vaccine or in the development of a diagnostic assay. In this study, patterns of nucleotide polymorphism were used to test the hypothesis that natural selection promotes diversity in different parts of SA G1 of Sarcocystis neurona. Nucleotide and amino acid sequence analysis of SA GI from multiple S. neurona isolates identified two alleles. Sequences were identical intra-allele and highly divergent inter- alleles. Also, phylogenetic reconstruction showed sequences clustering into two clades. Tajima 's and F u and Li '3 neutrality tests indicated that selection is more likely to be acting on SA GI . Moreover, a sliding window analysis based on the ratio of silent substitutions to amino acid replacements provided strong evidence that two short segments in the central and 3 ' domain of SA G1 have been under positive selection in the divergence of the two alleles, suggesting that it may be important for the evasion of host immune responses and would be a suitable target for vaccine deve10pment. 114 INTRODUCTION Sarcocystis neurona, the major causative agent of equine protozoal myeloencephalitis (EPM) in horses, is widespread throughout the New World (Dubey et al., 2001). EPM is a neurological disease that imposes a serious burden on the horse industry in the United States, for which there is no effective vaccine. Seroprevalence rates of S. neurona among horses can reach very high levels, such as 60% in Michigan (Rossano et al., 2003) and 89.2% in Oklahoma (Bentz et al., 2003). Even though a large percentage of horses are exposed to this parasite, the actual incidence of EPM in horses is very low ~1% (MacKay et al., 2000). The design of an effective vaccine is complicated by the fact that field isolates differ in immunogenicity (Mansfield et al., 2001; Marsh et al., 2001, 2002). Additionally, genetic information for the distinction of S. neurona isolates is lacking. S. neurona has an immunodominant surface antigen gene 1 (SA G1), which appears to occur as a single copy and encodes a major antigenic surface protein of about 29 kDa (Ellison et al., 2002). SAG] has been suggested as a diagnostic marker for phylogenetically related Sarcocystis species from different geographical areas (Hyun et al., 2003). The genetically closely related organism T oxoplasma gondii has an immunodominant major surface antigen P30, which has been extensively studied with a view to the development of a vaccine (Bonenfant et al., 2001; Letscher-Bru et aL,2003) Positive selection and sequence polymorphism are among the major factors that determine how the gene has evolved and, consequently, may be important for the development of a vaccine. Synonymous (silent) mutations are largely invisible to 115 natural selection, while nonsynonymous (amino acid-altering) mutations can be under strong selection pressure (Akashi, 1995). The rates for nonsynonymous (p3) and synonymous (pg) mutations were defined as the numbers of substitutions per site. Comparison of the nonsynonymous/synonymous substitution rate ratio (w=dN/dg) provides a powerful tool for understanding the effect of natural selection on molecular sequence evolution and can be used as an indicator of selection pressure at the protein level (Ohta, 1993). In general, amino acid sites in a protein are expected to be under different selection pressures and have different underlying wratios, where w=1 indicates neutral mutations, w<1 means purifying selection, and w>1 indicates positive selection. There is a paucity of published data on the extent of sequence polymorphism and evolution of the outer membrane protein genes, including SA GI of S. neurona. Therefore, the objectives of this study were to examine the molecular basis of genetic variation and adaptive molecular evolution of SA G1 of S. neurona and to identify the amino acid sites of this gene under diversifying selection. The main findings of this study are that: (1) parts of S. neurona SA GI have evolved under positive selection, and (2) SA GI has two alleles in the limited number of S. neurona isolates examined. 116 MATERIALS AND METHODS Nucleotide sequences. All SA GI sequences used in the study were obtained from GenBank. Nine SA G1 sequences of S. neurona (accession nos. AF401682, AF397896, AY032845, AY245695, AY170900, AF480854, AY170620, AF480853, AY245696) were used as in—groups. One T. gondii SAGI sequence (accession no. AY217784) was used as an out-group to root the phylogenetic tree(s). Sequence alignment and phylogenetic analyses. Nucleotide sequences were aligned using the multiple sequence alignment program Clustal X (Thompson et al., 1997) with the default parameters and checked by eye. Base frequencies and pairwise sequence divergences were obtained using the program PAUP. version 4.0b10 (Swofford, 2002). All insertion-deletion events were coded as missing data (“17") for the purpose of phylogenetic analysis. Phylogenetic trees were sought with both maximum parsimony (MP) and neighbor-joining (NJ) using MEGA (Molecular Evolutionary Genetics Analysis, version 1.01) software (Kumar et al., 1994). To find the shortest (MP) tree, a branch-and-bound search was performed with unknown initial upper bound and ‘furthest’ taxa addition sequence. NJ searches were performed to assess the influence of alternative phylogenetic algorithms on the tree topology using the Kimura 2-parameter (K2P) distance model of sequence evolution. Relative bootstrap support for nodes in resulting trees (both by MP and NJ) was evaluated using 1,000 pseudoreplicates (Felsenstein, 1985). The maximum-likelihood (ML) tree was obtained using a heuristic search in PAUP* without topological constraints in which the parameter values under the best-fit model were fixed and a NJ tree was used as a starting point for TBR branch swapping. The resulting tree 117 topology and new parameter estimates were used in a second round of branch swapping to provide the final ML tree. Bootstrap support for nodes in the ML tree was evaluated for 100 pseudoreplicates using TBR branch swapping on starting trees obtained by neighbor joining. Bayesian phylogenetic analysis implementing the GTR+ F model was conducted using the software program MrBayes version 3.0 (Huelsenbeck and Ronquist, 2001). Four (nchains=4) heated (temp=0.5) Markov chains were calculated simultaneously and sampled every 100 generations (samplefreq=100) for 2,000,000 generations. The first 2,001 (of 20,001) trees were discarded (bumin=2,001). The posterior probabilities of each branch were calculated by counting the occurrence in trees that were visited during the course of the Markov Chain Monte Carlo (MCMC) analysis. Statistical tests of neutrality and sequence polymorphism. The numbers of polymorphic sites in the nucleotide and amino acid sequences of each identified allele were computed using the programs PSFIND (version 2) and PAP IND (version 1.1), respectively. These statistics were then used to draw the similarity plot (Figure 1A, B) using Happlot program (version 1.1). To examine how the level of selective constraint varies along different parts of SA GI, nucleotide site differences were calculated by the method of Nei and Gojobori, (1986) and tabulated in a sliding- window analysis of 30 codons for the length of the protein-coding region of the gene (Figure 2) using the program PSWIN (version 1.1). The difference pijs is a measure of the degree of selective constraint: the more negative the value, the less the contribution of replacement substitutions and the greater the contribution of synonymous substitutions. The zero-difference line indicates selectively neutral 118 variation, where the per-site rates of synonymous and nonsynonymous substitutions are equal. A positive difference, where amino acid replacements exceed the silent substitutions, suggests the action of diversifying (positive) selection. The PSF IND, PAF IND, and PSWIN programs were written and kindly provided by Dr. T.S. Whittam (Microbial Evolution Laboratory, National Food Safety and Toxicology Center, Michigan State University, East Lansing, MI 48824, USA) The Approximate Likelihood Ratio (ARS) analysis software (Pond and Frost, Antiviral Research Center, University of California, San Diego; http://www.datamonkey.org) was used to evaluate selection pressures on SA GI gene. This software uses a maximum-likelihood approach with codon-based models to estimate the ratio (at) of dN, the rate of nonsynonymous substitutions per nonsynonymous site, to d5, the ratio of synonymous substitutions per synonymous site (Figure 5). To determine if the sequence diversity observed in SA G1 is a consequence of diversifying selection, which might suggest that it is a target of protective immune response, additional measures of DNA sequence polymorphism, such as the number of segregating (polymorphic) sites, number of haplotypes (NHap), haplotype diversity (HapDiv), nucleotide diversity (0), and F—statistics, were computed using the DnaSP program version 3.52 (Rozas and Rozas, 1999). The genetic polymorphism was also estimated by the statistic 7r, which is the average number of nucleotide differences per site between two sequences. Because selection may have played a role in the evolution of SA GI of S. neurona, and divergence between isolates (given the putative short divergence time), polymorphism data for each allele were tested by Tajima 119 (Tajima, 1989), and Fu and Li (Fu and Li, 1993) neutrality tests. Tajima ’s test compares two estimates of genetic diversity: the mutation parameter (0) and the difference between pairs of isolates (71'). Fu and Li '5 test is related to Tajima 's and compares the number of singletons to that expected by the neutral theory. It was performed without an out-group. Gaps and a few sites with unreliable information were coded as missing and ignored. In general, mutations occur at intermediate frequencies under positive selection. Thus, positive selection results in a small number of segregating sites relative to the average number of pairwise nucleotide differences between sequences (77), which is reflected by positive values of Tajima’s test (Tajima, 1989) and also results in decreased mutation rates in external branches of the genealogy, which is reflected by positive values of Fu and Li ’5 test statistics (Fu and Li, 1993). To investigate the role of intagenic recombination in SA GI sequence polymorphism, the degree of linkage disequilibrium (or nonrandom association between variants of different polymorphic sites) was estimated using the DnaSP program, with the following parameters: D, D', and R. Both the two-tailed Fisher '3 X . . . exact test and the 2-test were computed to detenmne whether the assocratrons between polymorphic sites were, or were not, significant. 120 RESULTS Nucleotide sequence diversity. Aligned sequences used for phylogeny reconstruction were 1,685 bp in length. Of these, 852 (51%) were variable and 208 (12%) were parsimony informative. Mean base proportions were 20.9% A, 22.9% T, 28.2% C, .X . . . and 27.9% G. 2-test of homogeneity of base frequenc1es across taxa showed sllght non- significant bias to the nucleotides G and C; 334.36 A, 366.65 T; 451.30 C, and 447.69 G; x2=23.766 (d.f.=24), P=0.4749. Sequences were clustered into two grOUpS based on the uncorrected pairwise sequence divergence (p) model, and the degree of divergence between the two groups was large but minimal within each group (Table 1). Nucleotide similarity analysis revealed that within each monophyletic group of S. neurona isolates, sequences of SAG] were 100% identical while sequence similarity between the two groups was 71%. The fact that sequences within each cluster were identical and very different between the two clusters clearly indicates that there are two SA GI alleles, which, hereafter, we refer to as H1 and H2. 121 Table 1. DNA sequence differences of SAG] of Sarcocystis neurona isolates based on pairwise comparisons using an uncorrected (“p") distance matrix. .-.-...- ... ._—v—‘ —— --.... -....- ..---- -. l : l3. Sn-mucat2l(AY245695) 10.000000001- l l4.UCDl {(AF397896) 0.00010000100001 l 35.1mm l(AF40l682) 100001000010.00010000i- l l 36.SN3 i(AYO32845) 1000010001(00000001100011- l g l l l 22. Sn-mucat2 l(AY170900) [0.00014 1 ...—3...." ...—....q llsolate iAccession no.l1 ,2 i3 l4 35 6 l7____l8 "l9 {1.CAT2 1(AF480854) [4 . l T” W l l l l l [. i7.MU-1 I(AF480853) [0.243l0.237|0.243l0.238l0.238l0.238 1 l8. Sn-MUl [(AY245696) [024210.241 [0.243102421024210243 {0.0001- 9. Sn-Mul §(AY170620) 10.255 [0393 1024010335 [0.402 [0.240|0.00910.000 (- Phylogenetic inferences. The evolutionary relationships between the two SA G1 alleles were investigated using a phylogenetic analysis. A branch-and-bound search with equally weighted data produced a single tree, which was rooted by the T. gondii sequence. The SA GI tree (Figure 1) indicates clearly that there are two well-defmed clusters: cluster I consisted of six nucleotide sequences representing three S. neurona isolates; the SN3 and UCDl isolates from horses from Panama and California, respectively, and the CAT2 isolate from a cat from Missouri, and cluster 11 included three SA GI sequences of the isolate Sn-MUl from a horse from Missouri. The SAG] genealogy was also estimated using the NJ method with the K2P model since more complex models gave substantially similar results (data not shown). The tree obtained by the NJ method had identical branching patterns and very close bootstrap support to the tree found by the MP method (Figurel). Bootstrap values for NJ and MP methods obtained from 1,000 replications were robust with 100% support at the basal node of each 122 clade. Maximum likelihood and Bayesian phylogenetic analyses of the nucleotide sequences also consistently yielded trees with almost identical topology (data not shown), indicating that the inferred phylogenetic relationships are robust to different methods of tree estimation. S. neurona (AF397896. Horse, California) S. neurona (AF 401682, Horse, California) 52/51 S. neurona (AY032845, Horse, Panama) 77/ 75 S. neurona (AY170900, Cat, Missouri) 100/100'3. neurona (AY245695, Cat, Missouri) S. neurona (AF480854, Cat, Missouri) S. neurona (AF480853, Horse, Missouri) 100/100I S. neurona (AY245696, Horse, Missouri) 99/96 S. neurona (AY170620. Horse. Missouri) T. gondii (AY217784) |———l 0.1 substitution Figure 1. Maximum parsimonious phylogenetic tree from analysis of 208 phylogenetically informative characters of SA G] showing relationships among the Sarcocystis neurona isolates (rooted to Toxoplasma gondii). The first of the two numbers at the nodes represent bootstrap resampling results based on maximum parsimony analysis (MP, % of 1,000 replicates). The second number represents the bootstrap support using neighbor-joining algorithm (NJ, % of 1,000 replicates). Numbers in parentheses after taxon names refer to GenBank accession nos., followed by the host and origin of isolate. Tree statistics are length (L)=290, consistency index (CI)=O.997, retention index (RI)=0.998 excluding uninforrnative characters 123 Evidence for positive selection of amino acid changes. The degree of nucleotide and amino acid polymorphism and divergence between H1 and H2 alleles was remarkable. Variations in sequences were distributed almost across the whole of SA GI (Figure 2A—B). Nucleotide diversity (0) was also remarkably high (0.16545). The ratio of nonsynonymous to synonymous mutations in the divergence of the whole gene between the H1 and H2 alleles was 0.73:0.05, a ratio that would be predicted for unconstrained, neutral variation. Three exceptions were noticed. First, the region encoding the signal peptide sequence (Figure 3) showed the least polymorphism; amino acid replacements were highly constrained in this part of the gene. Secondly, there were two large segments where the per-site nonsynonymous-substitution rate peaked upward (Figure 4). These peaked regions of the gene showed the most polymorphism as shown in Figure 2A, B. Thirdly, SA GI showed much less constraint on amino acid-altering mutations in particular regions among alleles. Of the whole open reading frame (980 bp) of SA GI, only two stretches, located in the central and the 3' regions of the gene, were identified where there was a positive difference between the nonsynonymous- and synonymous-substitution rates. The longest stretch with dN/ds>1 was ~51 codons in length and ran from position 114 to position 165. In this region, the rate of substitution per 100 sites (corrected for multiple hits) was dN=1922t16 and ds=165i21. The second region extended from codon 222 to codon 259. In this region, the rate of substitution was dN=2962t15 and ds=255:19. Thus, a moderately positive selection pressure is acting on these regions of SA GI . 124 Approximate Likelihood Ratio (ARS) analysis also identified a positive difference between the nonsynonymous- and synonymous-substitution rates in the same two areas located in the central and the 5’ regions of the gene (Figure 5). There were two haplotypes, and haplotype diversity (expected allele heterozygosity) was 0.667 (variance 0.04167, SD 0.204). The number of polymorphic (segregating) sites was 204. This dataset had a nucleotide diversity of 0.16545 ( variance 0.0025663, SD 0.05066). A number of statistical tests were used to determine departure from neutrality at the SA GI locus. Tajima '5 test is the most powerful for testing alternative hypotheses involving selective sweeps, population bottlenecks, and population subdivision (Simonsen et al., 1995). For the SA GI sequence data of S. neurona, all values of neutrality tests were significant, indicating an excess of intermediate frequency variants and balancing selection (i.e., rejected the neutral model of molecular evolution). Tajima 's D test was 2.33106 with a P<0.01. The Fu and Li statistics, D and F had values of 2.33108 and 2.50897, respectively, both with P<0.02. Fu ‘5 Fs statistic was 12.990 (P>0.001). No linkage disequilibrium tests were significant (P<0.333) and there was no evidence for any recombination events. 125 A 5-prime Intron 30.5% 40% 3-prime 24% Domains S. neurona I A F401 682) S. neurona J (AY170620) B l Domains I I S-prime Ell ll|||||||l|ll|l|llClll| 3-prime S. neurona (AF401682) - S. neurona (AY170620) I II I lllllllll ll Illllllllllllll 11111] [I Figure 2. Diversity in SA GI gene of Sarcocystis neurona. Plot of the variable nucleotide sites (A) and amino acid sites (B) in pairwise comparisons of the two alleles of SA G1 gene with percentage of the nucleotide polymorphism shown above the domains. General three-domain structure model of Sarcocystis neurona SA GI gene is drawn above with the intron region shaded in grey. The arrow marks the signal peptide sequence region. Each breakpoint in the domains denotes the location of an alignment gap. Each vertical line marks the location of a nucleotide or amino acid difference between the two alleles sequence. 126 Domain = S-prime S. neurona AF401682 ATGACGACGGCGGTGCTGCTGACGTTTCTGACACTCTGCTCCGCCAGAGTGTCCCTTGTG 60 S. neurona AY170620 ............ A ........... A . C .......................... AT . C . . . S. neurona AF401682 AGGGCCGGAGCGCCGCCTCAAGCAACGTGCGCCAATGGCGAAACGACTGTTACTAAGCTC 120 SneuronaAYl70620 .AC ..... G.G.G.C.G ...... G..A....AAG.G...C ......... C ...... A... S. neurona AF401682 GGCAGCTCTGGCGCACTACGAATCCACTGCCCAAATAATTTTCGACTC---GCGCCCCGG 180 S.neur0naAYl70620.AG.A.C ...... T ..... A.C.A..A ..... GGC.CGG.A..A...GAATC..G.T.CC S. neurona AF401682 GCTGGGAATGACGCCGGTCAGATGCAGGTCTATGCAACTGCGGTTGCTGAGAATCCTGTA 2 4 O S. neurona AY170620. .C.C.G.C.C.AA. .AAGGT .......... T.A. . . .C. . . .C. .TC.G. . . .G. . . .G S. neurona AF401682 AACATACGAGACGTCCTGCCCGGCGCATCTTACCTCTCTGTACAGAACGTCCCG ------ 300 S. neurona AY170620 GCGC.T.A. .GT ..... C ......... A ..... G.TCG. .CG.CA. .T.GAG.TGGAGCT S. neurona AF401682 ACCCTCACCGTCCCGCAATTGCCCGCCAAAGCTACGAGCGTCTTTTTTCACTGCCAGCAG 360 SneuronaAYl70620..A..G...A ....... GC ..... C.A..G...G....T..G...A....A ........ A S. neurona AF401682 CAA---CCCGACAACCAATGCTTCATCCAGGTAGAAGTAGCGCCGGCTCCGCGCCTAGGT 420 S. neurona AY170620 . . . GGA . AAC . GGGA . . G ......... G ..... C ........ GG . T .............. Domain = intron S. neurona AF401682 AGGTTTATATGCATTAGGAGTAGGTTCATATGCATTAGGATGCGTACTCGTGGCTCGAAT 480 S. neuronaAYl70620 .A. . .CTC ------------ . .T. .ATG. . . .TG.--.TTCC.T.CG.T. .A. . . .T. .. S. neurona AF401682 GTTCCACTGCAGCGCGGTGAGATTGGCGGCTAACCGGGTAAATGTGCGTCTTTTTTTGTT 540 S.neuronaAYl70620.GAG ......... T.C...C..C....A ....... C..GT.TC....ACT .......... S. neurona AF401682 ACGCAGGT S. neurona AY170620 . . .T. . .- Domain = 3-prime S. neurona AF401682 CCGAATACCTGCGCGGCGCTGCAGTCCACGATCGCCTTCGAAGTTCAACAAGCGAATGAA 610 S. neurona AY170620 .................. G . A ....... G ..... AT ..... GA . .AG . GC . CA ....... S. neurona AF401682 ACAGCAGTCTTCAGCTGCGGCGAGGGACTTGCTGTGTTCCCGCAAGGTAGCAAAGCGTTG 670 S. neurona AY170620 G ....... G ............. C ..... GG. . .CC . . .T ..................... S. neurona AF401682 GATGAAGCCTGCTCCAAAGAGCAGGCCCTACCCAGTGGCGCCGCTTTAGCTCCAAAGGAT 730 S. neurona AY170620 ..... T ........ G ........ AT .......... C. . . .TG .................. S. neurona AF401682 GGTGGG---CTCCACCTTGGTTTTCCTCAGCTTCCTCAGCAGGCTATGAAGATTTGCTAT 790 S.neuronaAYl70620.C...TTCGT....G ..... C ..... G ..... C ...... A.CC..C.A ..... A..T... S. neurona AF401682 ATTTGTACGAATGGTGGTGTGCAGGCAGAGGCGGCC---CAACGGTGTGAGGTTCGCATC 850 SneuronaAYl70620 ..... C...G.A ..... CCA..G..T...T ..... GGAT...A ....... A....AT... S. neurona AF401682 TCCGTCGCAGCGAACCCAGACGGAAGCGTTCCAGGGGCTAACGGAGCCGCCTCTCTAGGA 910 S.neuronaAYl70620 G....T....G.GC.GA..C....G..CC.A ..... C...C ..... G...T...G.G..C S. neurona AF401682 GCTGCCGCACGCAGCGCCTCTGCGTTAGGGTTGGCTCTCGTTGCAGGCGCTTTCTTGCAC 970 S.neuronaAYl70620 C....T ................. A....TCC ..... G ......... C..G ...... C..T S. neurona AF401682 TTTTGCTAA 980 S. neurona AY170620 ..... G . G . Figure 3. Nucleotide alignment of 2 SA G] sequences representing the 2 alleles of S. neurona SA G] that shows predicted cleavage site for a signal peptide (underlined) as identified by Hyun et al., (2003). Dots indicate sequence identity and dashes indicate alignment gaps. 127 Pn—Ps Proportion of sites (x100) _. e a a '3 o s a s s s Codon pos'tion Figure 4. Plot of the number of substitutions per 100 sites for synonymous (pg) and nonsynonymous (p3) sites between H1 and H2 SAG] alleles in a 30-codon sliding window. The difference (Ml—pg) is a measure of the level of selective constraint on various parts of the molecule. 128 3.3.3.338 Amm< mmm a ._3 8 22.3352 333— 533522 333308035 363 .3: 95300 mmUEODmZ 253.3: .3. anwwnmc>< $3. a .138 23:35.2 33— cmeozz 333330.535 333 $5 933.50 vaEODmE 3:93»: .3. 30V33~c>< Sm a .E B 223ng 3333 533532 mumxuocoam A83 .3; 25300 mmUEODmZ 36.33: .3. «9.33335. 3.. a .3 8 23223 383 53222 35283 25533.: .3 3330:). 22%: .3 «3.33382 333 2833 ..m s 223222 33.3. 53222 $3283 2253...; .Q EOE. 28%: .3 3933.854. 333 :83 ..3 “o 22.33533. 83. 533222 mamxoeoam 25333.33 .Q 50:2 56.33: .3. 3N¢33Nc>< 333. :83 .3 8 23:35.2 Son 233222 3863832 A3358 ..3 38: 22:2 3058: .3. 33333.5.3 mmm 28$ ..3 8 22.3532 ooom 59:22 3263202 3:53 .3 038: SE: 26.53: .3. «ovwn~c>< N33 :88 ._3 8 Seaman—2 333— 533522 3263232 A3353 ..3 85: 33:2 38.53: .3. nomwhmc>< mmm COOS ..3 3 20:35.2 333— 5.3522 32.53982 33536 ..3 03.5: .52 38.33: .3. 9.23-:— .e= :23339< 35 53> 25550 53:9.— uuctouom 5:29.: 3:39.— 335 33:: cozaaufiua 33.8— .3353 :8on 2: E 333353 328% 3.3.3.0833. 2: no mzflom .3 03.3.3. 143 $33282. .3222 2.2.2283 223233.... 22.22 .3 £332.85... 2228 2.823382 ANmmR©>