MOLECULAR DETECTION, GENETIC DIVERSITY AND PHYLOGENETIC ANALYSIS OF ANAPLASMA MARG/NALE INFECTING CATTLE IN SOUTH AFRICA By Awelani Mirinda Mutshembele Thesis submitted in fulfillment of the requirements for the degree Philosophiae Doctor in the Faculty of Natural and Agricultural Sciences, Department of Zoology and Entomology, University of the Free State Promoters: Dr. M.S Mtshali and Prof O.M.M Thekisoe December 2015 PROMOTERS Dr. Moses S. Mtshali (PhD) Research and Scientific Services Department National Zoological Gardens of South Africa P.O. Box 754 Pretoria 0001 Department of Zoology and Entomology, University of the Free State, QwaQwa Campus Private Bag X 13 Phuthaditjhaba 9866 Prof. Oriel M.M. Thekisoe (PhD) Unit for Environmental Sciences and Management North - West University, Potchefstroom Campus Private Bag x 6001 Potchefstroom 2520 ii DECLARATION I, the undersigned, hereby declare that the work contained in this thesis is my own original work and that I have not previously in it's entirely or in part submitted it at any university for a degree. I furthermore, cede copyright of the thesis in favour of the University of the Free State .. . . .. ... ~~· ··· · 0.9..l 0.~ . \ .?-9.! .~ ... AM. Mutshembele Date iii DEDICATION This thesis is dedicated to my parents my late father Itani Daniel Mutshembele may his soul rest in peace, and my mother Tshilidzi Mercy Mutshembele for their endless love, support , encouragement and tolerance on my career choice. iv ACKNOWLEDGEMENTS I thank Almighty God for the strength He gave me to do research under supervision of Dr. M. S. Mtshali and Prof 0 . M. M. Thekisoe, without their assistance and dedicated involvement in every step throughout the process, this study would have never been accomplished. I thank them very much for their support and understanding over these past four years. I also thank our collaborators Prof Jose de la Fuente and Dr. Alejandro Cabezas-Cruz for their endless support during my research period. I would also like to show gratitude to my colleagues and friends in the Veterinary Parasitology Programme for their immeasurable contributions and not forgett ing the soul mates Dr. Raksha Bhoora, Dr. Tracy Masebe, Zama Khumalo, Lesego Modibedi - Masango, Portia Motheo, Sabath Rathanya, Verdi Giqwa and Seipati Nku for their support and patience during difficult study times and "Sissy Khanyi" Prof Khanyisile Mbatha for her support and advice throughout the year, she was my pillar. I also thank Moeti Oriel Taioe for his assistance with phylogenetic analys is at the Department of Zoology and Entomology, University of Free State QwaQwa Campus and Dr. Steve Olorunju at the Biostatistics Unit, Medical Research Council of South Africa, Pretoria who kindly assisted me with the statistical analysis and was very patient with me. v I thank my brothers Dr. Theodore Mutshembele, Ndivhoni Mutshembele, my sister Mudzunga Mutshembele and my nieces (Mapitso and Malindi). Every time I was ready to quit, they did not let me to quit and I am forever grateful, this thesis stands as a testament to your unconditional love and encouragement, and Abel Ralinala for his love, patience, jokes and words of inspiration that gave me strength and kept me motivated. Getting through my dissertation required more than academic support, and I have many, many people to thank for listening to me and, at times, having to tolerate me over the past three years. I cannot begin to express my gratitude and appreciation for their friendsh ip. I also thank the Directors of Department of Agriculture and Veterinary Institute from Limpopo, Mpumalanga, North West, KwaZulu-Natal , Eastern Cape, Western Cape and Northern Cape (Prof. Nomfundo Mnisi, Ors Mlilo, Songelwayo Chisi, L. Mrwebi, Ronald Sinclair and Francis Sabi) with the assistance from the Veterinary Technicians (Toypat Mdlul i, Keneilwe Constance Kaotane, Jan Masethe Maime, Anil Suresh and Erwin Lucas) and the farmers who co-operated in sample collection . I acknowledge the University of the Free State and National Zoological Gardens of South Africa for availing their facilities during this study. This study is based on the research supported in part by the National Research Foundation of South Africa from the DST/NRF Professional Development Programme, grant made available to Dr. M.S. Mtshali; and the National Zoological Gardens of South Africa. Any opinion, find ing and conclusion or recommendation expressed in th is study is that of the author and the NRF does not accept any liabi lity in th is regard. vi TABLE OF CONTENTS CONTENTS PAGE TITLE PROMOTERS ii DECLARATION iii DEDICATION iv ACKNOWLEDGEMENTS v TABLE OF CONTENTS vii LIST OF FIGURES xi LIST OF TABLES xv APPENDICES xvi LIST OF ABBREVIATIONS xvii RESEARCH OUTPUTS xix ABSTRACT xxi 1 INTRODUCTION AND LITERATURE REVIEW 1.1 Historical background 1 1.2 Classification of Anaplasma marginale 1 1.3 Epidemiology 2 1.4 Distribution of ticks infesting cattle in South Africa 4 1.5 Pathogenesis of A. marginale 7 1.6 T ra nsm ission 9 1.7 Life cycle of A. marginale 10 1.8 Geographic distribution and economic importance of anaplasmosis 12 1.9 Diagnosis of anaplasmosis 14 1.9.1 Competetive enzyme -linked immunosorbent assay 16 1.9.2 Card agglutination test 16 vii 1.9.3 Complement fixation test 17 1.9.4 Indirect fluorescent antibody test 17 1.9.5 Polymerase chain reaction 17 1.10 Treatment, prevention and control strategies 19 1.11 Anaplasma marginale major surface proteins and their role in host-vector- 21 pathogen interactions 1.11 .1 The MSP1 complex 22 1.11 .2 Vector-pathogen relationships 26 1.11 .2.1 Anaplasma MSPs and vector-pathogen interactions 26 1.12 Phylogenetic relationships of geographic isolates of A. marginale 27 2 OBJECTIVES OF THE STUDY 2.1 Statement of the problem 29 2.2 Aim of the study 32 2.3 Objectives 32 3 MOLECULAR DETECTION OF ANAPLASMA MARG/NALE ISOLATES INFECTING CATTLE IN SOUTH AFRICA BY PCR TARGETING msp1a GENE 3.1 Introduction 33 3.2 Materials and methods 34 3.2.1 Study site and sample collection 34 3.2.2 DNA extraction 35 3.2.3 Polymerase chain reaction (PCR) 35 3.2.4 Statistical analysis 36 3.3 Results 37 3.3.1 Detection of A. margina/e infections 37 3.3.2 Molecular prevalence of A. margina/e isolates infecting cattle in South 37 viii African provinces 3.3.3 Statistical prevalence of A. marginale isolates infecting cattle in South 37 African provinces 3.4 Discussion 44 4. EPIDEMIOLOGY AND EVOLUTION OF THE GENETIC VARIABILITY OF ANAPLASMA MARG/NALE IN SOUTH AFRICA 4.1 Introduction 43 4.2 Materials and methods 45 4.2.1 Study site and sample collection 45 4.2.2 DNA extraction 49 4.2.3 A. marginale species-specific PCR 49 4.2.4 DNA sequencing of A. marginale msp 1a gene 49 4.2.5 Sequence analysis of A. marginale msp1a gene 50 4.2.6 Codon based phylogenetic analysis of tandem repeats 50 4.2.7 Amino acid variability, composition and genetic diversity index (GDI) 50 4.3 Results and discussion 52 4.3.1 Molecular evidence of A. marginale prevalence in South Africa 52 4.3.2 A. marginale prevalence and msp1 a genetic diversity 55 4.3.3 Evolution of msp1a genetic diversity 56 4.3.4 Amino acid variability and low variable msp1a peptides 64 5. STRUCTURAL AND PHYLOGENETIC ANALYSIS OF ANAPLASMA MARG/NALE INFECTING CATTLE IN SOUTH AFRICA USING msp1a AND msp4 GENES 5.1 Introduction 66 5.2 Materials and methods 67 5.2.1 DNA extraction 67 5.2.2 Set of primers used for amplification of msp 1a and msp4 genes 68 5.2.3 Anaplasma marginale species-specific PCR of the msp1a gene 69 5.2.4 Anaplasma marginale species-specific PCR of the msp4 gene 69 ix 5.2.5 DNA sequence alignment and phylogenetic analysis 72 5.3 Results 73 5.3.1 Phylogenetic analysis of A. marginale isolates using msp1a gene 74 5.3.1.1 MSP1 a sequence analysis 74 5.3.2 Phylogenetic analysis using msp4 gene sequences 79 5.3.3 Phylogenetic analysis of msp4 gene using neighbor-joining phylogenetic tree 79 5.3.4 Phylogenetic analysis of msp4 gene using maximum-likehood phylogenetic 83 tree 5.4 Discussion 86 5.4.1 Phylogenetic analysis of A. marginale using msp1a gene 87 5.4.2 Phylogenetic analysis of A. marginale using msp4 gene 90 6. GENERAL DISCUSSION, CONCLUSIONS AND RECOMMENDATIONS 6.1 General discussion 92 6.2 Conclusions 93 6.3 Recommendations 95 REFERENCES 96 x LIST OF FIGURES Figure 1: Bovine erythrocytes infected with A. marginale. (A) Inclusion bodies located 8 at the periphery of the erythrocyte in a stained blood smear. (B) Electron micrograph of A. margina/e inclusion that contains four organisms (Zivkovic, 2010). Figure 2: The life cycle of A. marginale. The cycle is modified from Kocan (1999), to 11 include the replication of the rickettsia in endothelial cells (Carreno et al. , 2007) and Rhipicephalus microp/us ticks present in Mexico (Rodriguez et al. , 2009). Figure 3: Gel image of 1% agarose gel electrophoresis of PCR amplification products 39 obtained from A. marginale isolates from Western Cape Province in South Africa using 1733F and 2973R primers (Lew et al. , 2002). Figure 4: Prevalence of A. marginale by province in South Africa. 40 Figure 5: Map of South African areas included in the study. Map of South Africa 47 showing the provinces that were included in the study (Limpopo, Mpumalanga, North West, Gauteng, KwaZulu-Natal , Eastern Cape, Western Cape, and Northern Cape). Endemic, epidemic, and A. marginale-free areas are coloured (data collected from de Waal, 2000). The main tick species involved in the transmission of A. marginale in the sampled areas are shown: Rhipicephalus microplus, R. decoloratus, and R. evertsi evertsi (data collected from de Waal, 2000). Figure 6: Newly reported sequences of MSP1 a tandem repeats. The one-letter amino 54 acid code was used to depict msp1a repeat sequences. Dots indicate identical amino acids, and gaps indicate deletion/insertions. The ID of each repeat form was assigned previously in Cabezas-Cruz et al. (2013). Tandem xi repeat A was used as a model for amino acid comparison. Figure 7: Correlation between msp1a genetic diversity and A. marginale prevalence in 59 South Africa. The prevalences of anaplasmosis in different South African provinces were plotted against the average of genetic diversity index (GDI). GDI . was calculated for each strain as the number of different tandem repeats divided by the total number of tandem repeats. A polynomial correlation was found between these 2 parameters with R2 = 0.7 6. Provinces with 100% prevalence (Gauteng, Eastern Cape, and Mpumalanga) are in the range of 0.82-0.87 GDI while regions having less than 100% prevalence are out of this range. Figure 8: Reconstruction of ancestral amino acid sequence and amino acid positions 63 under positive and negative selection. The reconstruction of the ancestral state (TR 4: Tandem repeat 4, sequence marked with *) of the new tandem repeats found in South Africa (Figure 2) was performed using 3 reconstruction methods, namely: joint, marginal, and sample (see 'Materials and methods' for details). Positions that evolve under negative (-) and positive (+) selection are shown (see Table 4). Amino acid at position 20 is indicated (arrow). The residues of the immunodominant B-cell epitope (Garcia-Garcia et al. , 2004) (first box) and the neutralization-sensitive epitope (Palmer et al. , 1987; Allred et al., 1990) (second box) are also shown. Figure 9: MSP1 a amino acid variability, composition, and low-variable peptides. The 65 figure shows the amino acid variability and composition among the tandem repeats in 3 different countries: Venezuela (A), USA (B), and South Africa (C). Venezuela shows a high proportion of variable/conserved sites and a high average of amino acid variability while South Africa and the USA show middle and lower values, respectively. Different colours in columns depict different biochemical properties in the amino acid composition: negative xii (green), positive (red), uncharged-polar (beige), and non-polar (blue); proportion of deleted positions is shown in yellow. Consensus sequences of low-variable peptides are shown (*) for the USA and South Africa. The region of the immunodominant B-cell epitope from A. marginale (Garcia-Garcia et al. , 2004) is boxed in the low-variable peptide from South Africa. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.) Figure 10: Gel image of 1% agarose gel electrophoresis of PCR amplification products 71 obtained from A. marginale isolates from Western Cape Province in South Africa using MSP45F (de la Fuente et a/. ,2007) and MSP45R (Oberle and Barbet, 1993). Lane M: GeneRuler™ I Kb DNA ladder, ready-to-use; Lane 1- 10 indicate positive results and lane 11 indicate negative control and lane 12 positive control. Figure 11: Proportional identity between tandem repeat in position R1 and R1 +n. 75 Figure 12: Maximum Likelihood phylogenetic tree based on the msp1a from the strains 78 present in KwaZulu-Natal and North West provinces. The tree demonstrates that the strain A. marginale KZN-K (red dot) falls in the cluster from North West sequences (blue circle). Bootstrap values are shown as percentage in the internal branch. Only bootstrap values higher than 50 per cent are shown. NJ phylogenetic analysis provided similar results. Figure 13a: Neighbor-joining phylogenetic tree of A. marginale msp4 gene sequences 81 from strains identified in cattle in South Africa. The phylogenetic tree was implemented in the MEGA5 (Tamura et al., 2011 ). Bootstrap analysis was conducted with 1000 replicates. The GenBank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. xiii Figure 13b: Neighbor-joining phylogenetic tree of A. marginale msp4 gene sequences 82 from strains identified in cattle in South Africa represented with a green dot and blue dot representing an outgroups (A ovis and A. phagocytophilum) including isolates from West, East and North Africa , North and South America, Europe, Asia and Australia. The phylogenetic tree was implemented in the MEGA5 (Tamura et al., 2011 ). Bootstrap analysis was conducted with 1000 replicates. The Gen Bank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. Only bootstrap values higher than 50% are shown. Figure 14a: Maximum-likelihood phylogenetic tree of A. marginale msp4 gene 84 sequences from strains identified in cattle in South Africa. The phylogenetic tree was implemented in the MEGA5 (Tamura et al., 2011 ). Bootstrap analysis was conducted with 1000 replicates. The Gen Bank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. Figure 14b: Maximum-likelihood phylogenetic tree of A. marginale msp4 gene 85 sequences from strains identified in cattle in South Africa represented with a green dot and blue dot representing an outgroup including isolates from West, East and North Africa, North and South America, Europe, Asia and Australia. The phylogenetic tree was implemented in the MEGA5 (Tamura et al. , 2011 ). Bootstrap analysis was conducted with 1000 replicates. The GenBank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. Only bootstrap values higher than 50% are shown. xiv LIST OF TABLES Table 1: Ticks and the pathogens they transmit in cattle in South African provinces 6 Table 2: Estimated costs of tick and tick-borne diseases to cattle production 13 Table 3: Oligonucleotide sequences of primers used in this study 36 Table 4: Sampling sites and their coordinates in South African provinces 48 Table 5: Observed prevalences of A. marginale in different provinces of South Africa 51 Table 6: Anaplasma marginale strains and msp1a tandem repeats organization 57 Table 7: Sites that evolved under positive and negative selection in the new tandem 62 repeats from South Africa Table 8: Oligonucleotide sequence of primers used in this study and their references 68 Table 9: Anaplasma marginale strains and putative 20 structures of MSP1a tandem 76 repeats Table 10: Differences in the A. margina/e msp4 nucleotide sequences of 80 LP,NW,GP,KZN,WC and MP, EC isolates of South Africa xv APPENDICES Appendix 1: Raw data used for statistical analysis of A. marginale in cattle in provinces of 132 South Africa Appendix 2: GenBank accession number of A. margina/e msp1a gene isolates and their 139 origin in South Africa Appendix 3: GenBank accession number of A. marginale msp4 gene isolates and their 142 origin in South Africa Appendix 4: Amplified sequences of Anaplasma spp. isolates, their origin and GenBank 146 accession numbers xvi LIST OF ABBREVIATIONS oc degree Celsius BLAST Basic Local Alignment Search Tool BLASTn Basic Local Alignment Search Tool for Nucleotide bp base pair CAT Card Agglutination Test cELISA Competitive enzyme-linked immunosorbent assay CF Complement fixation test DNA deoxyribonucleic acid EC Eastern Cape EDTA Ethylenediamine tetra -acetic acid GDI Genetic diversity index GP Gauteng Province IFA Indirect fluorescent antibody Kb Kilo base kDa Kilodalton KZN KwaZulu-Natal LP Limpopo Province MAb Monoclonal antibody mM millimolar ml milliliter MP Mpumalanga MSP Major Surface Protein NC Northern Cape NCBI National Center for Biotechnology Information ng nanogram ng/µI nanogram per microliter NW North -West PCR polymerase chain reaction TBD Tick-borne diseases xvii UV ultraviolet WC Western Cape µg microgram(s) µm micrometer µM micromolar xviii RESEARCH OUTPUTS Publications Awelani M. Mutshembele. Alejandro Cabezas-Cruz, Moses S. Mtshali, Oriel , M. M. Thekisoe, Ruth C. Galindo and Jose de la Fuente. 2014. Epidemiology and evolution of the genetic variability of Anaplasma marginale in South Africa. T icks and Tick -borne diseases, Volume 5, Issue 6, 624-631 . Conferences Awelani M. Mutshembele, A. Cabezas-Cruz, M. S. Mtshali, 0 . M. M. Thekisoe, R. C. Galindo, J de la Fuente. Epidemiology and evolution of the genetic variability of Anaplasma marginale in South Africa. 81h International Ticks and Tick-Borne Pathogens and Biennial Social Tropical Veterinary Medicine (TTP-STVM) Conference. Cape Town, South Africa. 24-29 August 2014. Awelani M. Mutshembele, M. S. Mtshali, 0 . M. M. Thekisoe, R. C. Galindo, A. Cabezas-Cruz, J de la Fuente. Molecular prevalence, genetic diversity and phylogenetic analysis of Anaplasma marginale isolates in cattle in South Africa. 42"d PARSA Conference hosted by the North West University, Stonehenge in Africa, Parys, Free State, 22-24 September 2013. Seminars and Symposia Awelani M. Mutshembele, M.S. Mtshali and 0 . M. Thekisoe. Molecular detection, genetic diversity and phylogenetic analysis of Anaplasma marginale isolates in cattle in South Africa. Postgraduate seminar day, University of the Free State- QwaQwa campus. 13-14 November 2013. Awelani M. Mutshembele, A. Cabezas-Cruz, M. S. Mtshali, 0 . M. M. Thekisoe, R. C. Galindo, J de la Fuente. Structura l and epidemiological analysis of Anaplasma marginale using the major surface protein 1a in South Africa. 4 th National Zoological Gardens of South Africa (NZG) Research Symposium, November 2013. xix Awelani M. Mutshembele, Moses S. Mtshali, Oriel M. M. Thekisoe. The genetic diversity and phylogenetic analysis of Anap/asma marginale strains in cattle in South Africa. The Zoology Department Seminar, Faculty of Zoology and Entomology, University of the Free State, Bloemfontein Campus, November 2012. Awelani M. Mutshembele, Moses S. Mtshali, Oriel M. M. Thekisoe. The genetic diversity and phylogenetic analysis of Anaplasma marginale strains in cattle in South Africa. 3rd National Zoological Gardens of South Africa (NZG) Research Symposium, November 2012. Awelani M. Mutshembele, Moses S. Mtshali, Oriel M. M. Thekisoe. Analysis of phylogeographic relationships of Anaplasma marginale from South African isolates using msp4 gene sequences. The Zoology Department Seminar, Faculty of Zoology and Entomology, QwaQwa campus. November 2011 . xx ABSTRACT Bovine anaplasmosis caused by Anap/asma marginale is endemic in South Africa. This endemicity is due to presence of tick vectors that transmit A. marginale the causal agent of the disease and the high seroprevalence in Limpopo, Free State and North West provinces. To date, the genetic diversity of A. marginale isolates infecting cattle in all South African provinces, except Free State, are generally unknown. Recently, vaccines based on the A. marginale major surface protein 1a (MSP1 a) has been proposed as a strategy for controlling bovine anaplasmosis. However, characterization of genetic diversities of the A. marginale isolates in these regions is still needed before this protein can be used for vaccine development. Therefore, the aim of this study was to determine the prevalence, genetic diversity and phylogenetic relationship of A. marginale infecting cattle in all South African provinces except the Free State. A total of 280 whole blood samples were collected from cattle in all provinces with exception of the Free State. Twenty six districts and municipalities were included in this sampling. Anaplasma marginale genomic DNA was then extracted from the blood sample using ZR Genomic DNA 1M Tissue Miniprep (Zymo Research, CA, USA). A polymerase chain reaction (PCR) was done with primers targeting msp1a and msp4 genes and the PCR products were sequenced using genetic analyser (ABI , Life technologies, CA, USA). The generated sequences were analysed by bioinformatics and their phylogeny as well as genetic diversity index (GDI) was determined based on the sequences of msp1a and msp4 genes. xxi Overall , the prevalence of A. marginale infection in cattle was 76% in all provinces except for Northern Cape Province where the prevalence was zero. The prevalence per province was as follows: Eastern Cape 19.1% , Gauteng 9.6%, KwaZulu-Natal 23.0%, Limpopo 15.3%, Mpumalanga 10.1%, North West 12.4% and Western Cape 10.5%. The msp1 a revealed genetic variability with regions of different types of tandem repeats. Some repeats were conserved amongst the A. marginale strains and revealed low variable peptides in the MSP1 a tandem repeats. A polynomial correlation (R2=0.7 6) was observed between the GDI and anaplasmosis prevalence per province. Interestingly, provinces with the highest prevalence were not the ones with highest or lowest GDI. The analysis of msp4 gene sequences, which provided evolutionary information about geographically distinct A. marginale strains, was used in the present study for phylogenetic analysis of samples from Limpopo (LP), Mpumalanga (MP), North West (NW), Gauteng (GP), KwaZulu-Natal (KZN), Eastern Cape (EC) and Western Cape (WC) provinces of South Africa. Two clades were observed which consisted of first clade (LP, NW, GP, KZN and WC) and second clade (MP and EC) isolates. In addition when DNA sequence variation of msp4 gene was analysed in combination with isolates from other countries outside South Africa , important phylogeographic information was observed. The South African strains had 100% identity with isolates from Kenya, Zimbabwe and Australia. Good representation of the Southern and Northern Hemispheres was observed and demonstrated that the msp4 gene was a good phylogeographic marker. These results indicated that A. marginale is widespread in South Africa, and suggested that the analysis of msp1a and msp4 gene sequences provided an understanding of the phylogeny and epidemiology of A. marginale in South Africa. xx ii CHAPTER 1 INTRODUCTION AND LITERATURE REVIEW 1.1 Historical background The genus Anaplasma was established in 1910 by Sir Arnold Theiler who first described the "marginal points" in stained erythrocytes of sick cattle as the causative agents of a specific disease (Theiler, 191 O; 1911 ). These erythrocytic inclusions had been seen frequently in red blood cells of anemic cattle often in those suffering from babesiosis (piroplasmosis). The acute phase of the disease is characterized by weight loss, fever, abortion, lowered milk production and often death (Kuttler, 1984). 1.2 Classification of Anaplasma marginale Bovine anaplasmosis is a tick-borne rickettsial disease caused by the hemoparasite Anaplasma marginale (Aubry and Geale, 2011 ). Anaplasma marginale is classified within the Order Rickettsiales which was recently reorganized into two fam ilies, Anaplasmataceae and Rickettsiaceae based on genetic analysis of 16S rRNA, groELS and surface protein genes (Dumler et al. , 2001 ). Members of the family Anaplasmataceae are obligate intracellular organisms found exclusively within membrane-bound vacuoles in the host cell cytoplasm. The genus Anaplasma includes three species that infect ruminants, namely A. marginale, A. centrale and A. ovis (Dumler et al., 2001 ), also included within this genus is A. phagocytophilum, agent of human granulytic ehrlichiosis [HEG], A. bovis and A. platys (Kocan et al., 2010). Anaplasma marginale is a gram- negative rickettsia which is endemic in tropical and subtropical areas throughout the world (Decaro et al. , 2008; Howden and Geale, 201 O; Kocan et al. , 201 O; Aubry and Geale, 2011 ). Bovine anaplasmosis is a major constra int to cattle production in many countries (Kocan et al., 2010). 1 In order to persist in nature, A. marginale infects the mammalian host which usually remains persistently infected (Kieser et al., 1990; Eriks et al., 1993) serving as a reservoir for infection of ticks or mechanical transmission by the transfer of blood from infected to susceptible cattle (Kocan et al. , 1992; Ge et al., 1996; Futse et al., 2003). Taxonomic Classification of Anaplasma infecting ruminants (Dumler et al. , 2001 ): Kingdom: Bacteria Phylum: Proteobacteria Class: Alphaproteobacteria Order: Rickettsiales Family: Anaplasmataceae Genus: Anap lasma Species: A. marginale A. centrale A. bovis A. ovis A. phagocytophilum A. platys 1.3 Epidemiology Bovine anaplasmosis occurs in tropical and subtropical regions of the world and the disease is a major constraint to cattle production in many countries. Anaplasma marginale, in contrast to A. phagocytophilum, is quite host specific, infecting only ruminants and causing disease primarily in cattle. In the U.S.A. it is enzootic throughout the southern Atlantic states, Gulf Coast states, and several of the Midwestern and Western states (Kocan et al., 2010). 2 However, it has been reported in almost every state in the U.S.A. This increasingly wide distribution likely resulted from transport of carrier cattle with subsequent mechanical or biological transmission from asymptomatic persistently infected cattle to susceptible ones (Kocan et al. , 2010). It is also endemic in Mexico, Central and South America, as well as in the Caribbean Islands. It is enzootic in all Latin American countries, with the exception of desert areas and certain mountain ranges, like the Andes. The seroprevalence rates of A. marginale vary widely among countries in the America and the variability of these rates contributes to the development of geographically stable enzootic regions (Kocan et al. , 2010). In Europe, A. marginale is found mainly in Mediterranean countries with infections having been described in cattle and assorted wildlife species. It is also endemic in regions of Asia and Africa. The distribution of anaplasmosis may be expected to continue to change, in part as a result of global warming, which may influence the movement of the tick hosts. An example of the validity of such a prediction is a confirmed diagnosis of anaplasmosis in a bison herd in Saskatchewan, Canada, during the summer of 2000. The first reported outbreak of anaplasmosis in Canada occurred in 1971 , but this outbreak resulted from mechanical transmission from imported carrier cattle to local ones (Kocan et al. , 2010). In South Africa only one study has been conducted to genetically characterize the geographical strains (Mtshali et al., 2007). The results presented by the latter authors indicated the presence of a common genotype between South African, American and European strains of A. marginale. However the study focused only on blood samples of cattle collected from Free State province, so there is a need of further documentation of prevalence and genetic diversity of A. marginale strains in cattle of South African origin in order to design epidemiological and control strategies. 3 1.4 Distribution of ticks infesting cattle in South Africa Several species of ticks are found in different geographical areas in South Africa. The type of vegetation, humidity and temperature influences their distribution. Studies in the eastern Free State revealed that ticks affect mainly cattle of small farmers and the main tick species in the area are Rhipicephalus decoloratus (53.1% ), R. evertsi evertsi (44.7%), R. fol/is (1 .0%), R. gertrudae (0.7%) and R. warburtoni (0.4%). In the south west region of the same province, Fourie and Horak (1991) observed that Amblyomma marmoreum, Hyalomma marginatum rufipes and H. truncatum were the predominant species. A second study by Fourie et al. (1996) in the south west region revealed that lxodes rubicundus and H. m. rufipes were the most prevalent tick species (Marufu, 2008). In KwaZulu-Natal, Baker et al. (1989) observed that R. decoloratus, R. appendiculatus and R. evertsi evertsi were the most prevalent tick species of cattle raised on commercial farms. H. m. rufipes , R. appendiculatus and R. evertsi evertsi were the most numerous species in the Limpopo Province also present were R. microplus, R. decoloratus, A. hebraeum, H. truncatum and R. simus. Bryson et al. (2002) noted that the adults of A. hebraeum, R. appendiculatus and R. evertsi evertsi were the most numerous tick species in North West Province. In Mpumalanga, R. decoloratus constituted more than 75% of the total tick population. In a five year survey conducted in the Eastern Cape Province, R. decoloratus, A. hebraeum, R. appendiculatus and R. evertsi evertsi were found to be the most common tick species infesting cattle (Marufu, 2008). 4 In contrast to these early findings, Horak (1999) observed A. hebraeum, Haemaphysalis silacea, R. appendiculatus and R. glabroscutatum to be most prevalent tick species on yearling commercial cattle on Valley Bushveld. Muchenje et al. (2008) in an on station study comparing tick loads on Nguni, Angus and Bonsmara steers grazing on sweet rangeland revealed that R. decoloratus, A. hebraeum, R. evertsi evertsi and Hyalomma species were the most common tick infestations. With the exception of Hyalomma species, the same tick species composition observed by Muchenje et al. (2008), was found to infest cattle and goats in the communal areas of the Eastern Cape (Nyangiwe and Horak, 2007). Many studies have focused on cattle in the commercial farming system and it is apparent that cattle management and tick control in this farming sector will differ considerably to that in communal farming areas (Bryson et al., 2002). Few studies have focused on cattle kept by resource-poor farmers in communal areas. These studies however, did not compare the prevalence of the parasites in different breeds kept under communal farmer management and across different rangelands types. Other studies have focused on comparing tick loads in indigenous and exotic beef breeds under controlled conditions (Norval et al. , 1996; Muchenje et al. , 2008). No studies have focused on the comparison of tick loads in the indigenous and nondescript cattle under communal grazing management. Information on the tick loads of cattle can be used in conjunction with sere-diagnostic methods to estimate and compare the level of resistance of different cattle breeds to ticks (Wambura et al., 1998; Mattioli et al., 2000). Tick distribution and occurrence differs with geographic distribution and vegetation type (Mtshali et al. , 2004). No efforts have been made to compare tick loads in cattle on sweet and sour rangelands. Sour rangeland occurs in areas with high water supply and denser vegetation cover (Ellery et al., 1995) and it is more likely to have higher prevalence of ticks than the sweet rangeland which occurs in areas with low water supply and sparse vegetation cover. Comparing the prevalence of ticks in different rangeland types assists policy makers to design appropriate control programmes for each particular rangeland type (Marufu, 2008). 5 The most common tick species and the disease that they transmit to cattle in South Africa are shown in Table1 (Marufu, 2008). Table 1: Ticks and the pathogens they transmit to cattle in South African provinces Tick species Pathogens transmitted Geographical distribution (Provinces) Rhipicephalus (Boophilus) Anaplasma marginale, Limpopo, Mpumalanga, spp Gauteng, North West, Babesia bigemina, KwaZulu-Natal , Eastern B. bovis and Northern Free State, Eastern Cape and Coastal ~rip ~ Sou~ern and South-western Cape R. appendiculatus Theileria parva Limpopo, Mpumalanga, KwaZulu-Natal, Western T. annulata Cape T. taurotragi T. mutans T. velifera R. evertsi evertsi A. marginale Eastern half of Southern Africa, South-western Cape Hylomma spp A. marginale Karoo, Southern Africa except coastal Eastern Cape, KwaZulu-Natal, Eastern Southern and South-western Cape Amblyomma hebraeum Ehrlichia (Cowdria) Limpopo, Mpumalanga, ruminantium KwaZulu-Natal and Eastern and Southern T. mutans South-western Cape Sources: Horak et al. , 1991 ; Walker, 1991 ; Norval et al., 1992; Coetzer et al. , 1994. 6 1.5 Pathogenesis of A. marginale Anaplasma marginale is known to infect only mature, circulating erythrocytes of domestic and wild ruminants in vivo (Figure 1A). However, recently in vitro and in vivo studies indicated that it also infects endothelial cells, which may have implications for both pathogenesis and immune mechanisms. A. marginale enters erythrocytes by endocytosis and resides in the membrane-bound vacuole, where it divides by binary fission . The membrane vacuole is derived from erythrocyte membrane and contains 4 to 8 organisms (Figure 1B ). In the acute infection as much as 70% of erythrocytes may become infected (Zivkovic, 2010). During high rickettsiemias (bacteremia) multiple infections of individual erythrocytes are common. The incubation period varies with the infective dose and ranges from 7 to 60 days, with an average of 28 days. A. marginale leaves the host cell without disrupting it. Erythrocytes that are physically or chemically altered during the course of the disease are recognized by bovine reticulo-endothelial cells and phagocytized, which will result in the development of mild to severe anemia and icterus, without hemoglobinemia and hemoglobinuria (Zivkovic, 2010). The acute phase of the disease may also include symptoms such as high fever, dramatic weight loss, abortion, lethargy and often death in animals older than 2 years. Calves less than one year of age develop a relatively mild form of disease. Importantly, cattle that survive acute disease develop a lifelong persistent infection and serve as reservoirs for transmission to new susceptible hosts. Persistence is characterized by sequential rickettsemic cycles, occurring at approximately 5 week intervals, with peaks at 106 bacteria /ml of blood followed by a rapid decline when rickettsemia is controlled by a specific immune response. Persistently infected cattle have a lifelong immunity and are resistant to clinical onset of the disease on challenge exposure (Zivkovic, 2010). 7 Figure 1: Bovine erythrocytes infected with A. marginale. (A) Inclusion bodies located at the periphery of the erythrocyte in a stained blood smear. (B) Electron micrograph of A. marginale inclusion that contains four organisms (Zivkovic, 2010). 8 1.6 Transmission Transmission of A. marginale can be both mechanical by biting flies or blood- contaminated fomites and biologically by ticks. Mechanical transmission frequently occurs via blood-contaminated fomites , including needles, dehorning saws, nose tongs, tattooing instruments, ear-tagging devices, and castration instruments. Mechanical transmission by arthropods has been reported for bloodsucking diptera of the genera Tabanus, Stomoxys, and mosquitoes. This form of mechanical transmission is considered to be the major route of dissemination of A. marginale in areas of Central and South America and Africa where tick vectors do not occur and where Rhipicephalus (Boophilus) microplus, the tropical cattle tick, does not appear to be a biological vector of A. marginale. In areas of the United States where geographic isolates of A. marginale are not infective for ticks or where ticks have been eradicated by fire ants, mechanical transmission appears to be the major mode of A. marginale transmission (reviewed by Kocan et al. , 2010). In addition to mechanical and biological transmission, A. marginale can be transmitted from cow to calf transplacentally during gestation. For example, a 15.6% prevalence rate of in utero transmission of Anaplasma infections was reported in South Africa Transplacental transmission of anaplasmosis may therefore contribute to the epidemiology of this disease in some regions. This obligate intracellular pathogen can be transmitted biological by ticks, and approximately 20 species of ticks have been shown experimentally to transmit A. marginale worldwide. Tick transmission can occur from stage to stage (transstadial) or within a stage (intrastadial), while transovarial transmission from one tick generation to the next does not appear to occur. lnterstadial transmission of A. marginale has been demonstrated by the three-host ticks Dermacentor andersoni and D. variabilis in the United States and by R. simus in South Africa. The one-host tick R. annulatus transmits A. marginale in Israel, Central America, South America, and Mexico (reviewed by Kocan et al., 2010). 9 lntrastadial transmission of A. marginale is effected by male ticks. Recent studies have demonstrated that male Dermacentor ticks may play an important role in the biological transmission of A. marginale because they become persistently infected with A. marginale and can transmit the parasite repeatedly when they transfer among cattle. Male ticks therefore also serve as reservoirs of A. marginale along with persistently infected cattle. Transmission of A. marginale by male ticks may be an important mechanism of transmission by one-host ticks, including Rhipicepha/us (Boophilus) spp. and D. albipictus. However, it was shown recently that the co-feeding of adult Dermacentor spp. does not appear to influence the dynamics of A. marginale transmission (reviewed by Kocan et al., 2010). 1.7 Life cycle of A. marginale The life cycle of A. marginale in ticks is complex and well-coordinated with the tick feeding cycle (Figure 2). Infected erythrocytes are ingested by ticks with a blood meal and the first sites of infection are gut and malpighian tubule cells. During the subsequent feeding many other tissues, including salivary glands, become infected from where A. marginale can be transmitted to the vertebrate host. At each site of infection two stages of A. marginale occur within a membrane bound vacuole in the tick cell cytoplasm (Zivkovic, 2010). The first form seen within A. marginale colonies is the reticulated (vegetative) form, which divides by binary fission and results in formation of large colonies containing hundreds of organisms. The reticulated forms are then transformed into dense forms, which are the infective form and can survive for a short time outside of cells. Cattle become infected when the dense form is transmitted during tick feeding via the salivary glands. Ticks are able to acquire infection after feeding on persistently infected animals with a very low level of rickettsemia. Moreover, once ticks acquire the infection the biological replication of the organism with in the ticks makes up for the initial low infective dose (Zivkovic, 2010). 10 ee In ction o dense forms A marginale in the bovine Endothelial Cells Rhipicephalus Erythrocytes spp Replication in sevenl • tissues within the tick •• • .. .. .. • I • Feed acquisition of infected erythrocvtes / Rhipicephalus spp rigure 2: The life cycle of A marginale. The cycle is modified from Kocan (1999), to include the replication l~ the rickettsia in endothelial cells (Carreno et al., 2007) and Rhipicephalus microplus ticks present in rexico (Rodriguez et al., 2009). I 11 1.8 Geographic distribution and economic importance of anaplasmosis The distribution of anaplasmosis may be expected to continue to change in part as a result of global warming, which may influence the movement of the tick hosts (Jonsson and Reid , 2000. Concerns regarding the transmission of infectious agents between wildlife and domestic livestock are increasing especially in areas where free-ranging wildlife and cattle share common grazing grounds (Chomel et al., 1994). Bovine anaplasmosis is endemic almost anywhere the world from the Far East to Australia, Africa, Europe and the Americas (Wen et al. , 2002; Ziam and Benaouf, 2004; Naranjo et al. , 2006; Videtto et al., 2006; Stevens et al. , 2007). Bovine anaplasmosis causes important economic loss in most countries, mainly due to the high morbidity and mortality in susceptible cattle herds (Marufu, 2014). Losses due to anaplasmosis are measured through several parameters: low weight gain, reduction in milk production, abortion, the cost of anaplasmosis treatments, and mortality. However, few controlled studies have been carried out to determine the exact annual loss caused by anaplasmosis. The most important economic constraint of anaplasmosis to cattle production in the tropics is on public or private programs for genetic improvement of cattle. Imported Bos Taurus cattle brought from temperate nations to the tropics for breed improvement are highly susceptible to tick-borne diseases (TBD), and often do not survive to become part of planned production programs. This constraint is a notable reality for programs for the improvement of cattle in most Latin American countries (Marufu, 2014). The lack of accurate data on the epidemiology of ticks and TBD makes it difficult to determine their impact. The complexity of determining the direct and indirect economic impact of ticks and TBD, and their control is reflected in the fact that only rough estimates are available for the cost of some of the components. Table 2 shows the estimated costs of ticks and TBD to cattle production in different countries. Although a fairly crude estimate, these values may help to comprehend the importance of ticks and TBD in cattle. These estimates, however, expose the need for more studies on the determination of the economic impact of ticks and TBD in the cattle industry especially in the developing world (Marufu, 2014). 12 Table 2: Estimated costs of tick and tick-borne diseases to cattle production Country Costs (US$) Reference Global 13-18 billion de Castro (1997) Southern 31.6 million Minjauw et al., (1998) Africa Australia 4.09 million Jonsson et al., (2001) India 498.7 million Minjauw and Mcleod (2003) Australia 170 - 200 million Playford (2005), Sackett et al., (2006) Brazil 800 million Martinez et al., (2006) Losses are partly due to the direct effects of ticks on cattle, such as damaged hides and skins, anaemia, reduced body weight gains and milk yield, tick toxicoses and mortalities (Gates and Wescott, 2000; Turton, 2001 ; Mtshali et al. , 2004; Kaufman et al. , 2006). The damage caused by tick bites also diminishes the value of skins and hides for the manufacture of leather. Ticks with long mouth parts may induce abscesses because of secondary bacterial infections. Depending on the site of infestation, these abscesses can lead to lameness or mastitis resulting in the drop in milk production and subsequent increase in calf mortalities (Marufu, 2014). A large component of the economic cost of ticks in cattle is the appl ication of control measures to reduce infestations (de Castro, 1997; Porto Neto et al. , 2011 ). Conventional tick control is based on the appl ication of acaricides. The practice of intensive tick control spread rapidly throughout Africa following the introduction of imported cattle breeds and, in South Africa, it was enforced through legislation. There are few global reports on the costs involved in tick control and TBD treatments (Marufu, 2014). 13 Financial analysis revealed that spraying cattle with acaricide twice a week yields a return of 244% and maximised financial benefits to the farmer (Mukhebi et al. , 1989). Jonsson et al., (2001 ), however, estimated the total costs of tick control to contribute up to 49% of the total costs of ticks and TBD on the dairy industry in Australia. Expenditures for tick control were estimated at US$ 8.43, 13.62 and 21 .09 per animal per year for plunge dipping, hand spraying and pour-on, respectively (D'haese et al., 1999). The mean annual cost of ticks and TBD control per animal in pastoral and ranch herds was estimated to be US$4.54 (Ocaido et al., 2009). The development of new acaricides is also a lengthy and costly process leading to increasing cost of the newer products. Regular dipping has also led to the loss of resistance to ticks and enzootic stability to TBD. Significant losses also arise indirectly due to the important role of ticks in the transmission of TBD (Marufu, 2014). 1.9 Diagnosis of anaplasmosis The diagnostic assays (excluding cl inical findings) used to identify A. marginale can be classified as microscopic, serologic assays such as Indirect fluorescent antibody (IFA), complement fixation (CF) test, capillary, agglutination assay, card agglutination test (CAT), indirect fluorescent antibody (IFA) test, as well as various enzyme linked immunosorbent assays (ELISA) such as a cELISA, indirect ELISA and dot ELISA. The preferred serological tests are cELISA and CAT and molecular diagnostic assay using polymerase chain reaction (PCR) of whole blood samples (revised in Aubrey and Geale, 2011 ). However, the use of microscopy requires careful examination in cases of low level of parasitemia and A. marginale appear as dense, rounded and deeply stained intraerythrocytic bodies, approximately 0.3- 1.0 µm in diameter. Most of these bodies are located on or near the margin of the erythrocyte. This feature distinguishes A. marginale from A. centrale, as in the latter most of the organisms have a more central location in the erythrocyte. It can be difficult to differentiate A. marginale from A. centrale in a stained smear, particularly with low levels of rickettsaemia (OIE Terrestrial Manual 2012). 14 Anaplasma infections usually persist for the life of the animal. However, except for occasional small recrudescences, Anaplasma cannot readily be detected in blood smears after acute rickettsaemia. Thus, a number of serological tests have been developed with the aim of detecting persistently infected animals. A feature of the serological diagnosis of anaplasmosis is the highly variable results with regard to both sensitivity and specificity reported for many of the tests from different laboratories. This is due at least in part to inadequate evaluation of the tests using significant numbers of known positive and negative animals. Importantly, the capacity of several assays to detect known infections of long-standing duration has been inadequately addressed. An exception is cELISA, which has been validated using true positive and negative animals defined by nested PCR (Torioni De Echaide et al. , 1998), and the card agglutination assay, for which relative sensitivity and specificity in comparison with the cELISA has been evaluated (Molloy et al., 1999). It should be noted that there is a high degree of cross-reactivity between A. marginale and A. centrale, as well as cross-reactivity with both A, phagocytophilum and Ehrlichia spp. in serological tests (Dreher et al. , 2005; Al-Adhami et al., 2011 ). While the infecting species can sometimes be identified using antigens from homologous and heterologous species, equivocal results are obtained on many occasions (OIE Terrestrial Manual 201 2). 1.9.1 Competetive enzyme -linked immunosorbent assay A cELISA using a recombinant antigen termed rMSP5 and MSP5-specific monoclonal antibody (MAb) have proven very sensitive and specific for detection of Anaplasma- infected animals (Hofmann-Lehmann et al., 2004; Stik et al. , 2007; Reinbold et al. , 2010). All A. marginale strains tested, along with A. ovis and A. centrale , express the MSP5 antigen and induce antibodies against the immunodominant epitope recognised by the MSP5-specific MAb. A recent report suggests that antibodies from cattle experimentally infected with A. phagocytophi/um will test positive in the cELISA (Dreher et al. , 2005). 15 However, in another study no cross-reactivity could be demonstrated, and the MAb used in the assay did not react with A. phagocytophilum MSP5 in direct binding assays (Stik et al., 2007). Recently, cross reactivity was demonstrated between A. marginale and Ehrlichia spp, in naturally and experimentally infected cattle (Al- Adhami et al., 2011 ). Earlier studies had shown that the cELISA was 100% specific using 261 known negative sera from a non-endemic region , detecting acutely infected cattle as early as 16 days after experimental tick or blood inoculation, and was demonstrated to detect cattle that have been experimentally infected as long as 6 years previously (Knowles et al. , 1996). In detecting persistently infected cattle from an anaplasmosis-endemic reg ion that were defined as true positive or negative using a nested PCR procedure, the rMSP5 cELISA had a sensitivity of 96% and a specificity of 95% (Torioni De Echaide et al., 1998). 1.9.2 Card agglutination test The advantages of the card agglutination test (CAT) are that it is sensitive, may be undertaken either in the laboratory or in the field , and gives a result within a few minutes. Nonspecific reactions may be a problem, and subjectivity in interpreting assay reactions can result in variability in test interpretation. In addition, the CAT antigen, which is a suspension of A. marginale particles, can be difficult to prepare and can vary from batch to batch and laboratory to laboratory. Splenectomised calves are infected by intravenous inoculation with blood containing Anaplasma-infected erythrocytes. When the rickettsaemia exceeds 50%, the animal is exsanguinated, the infected erythrocytes are washed , lysed, and the erythrocyte ghosts and Anaplasma particles are pelleted . The pellets are sonicated, washed, and then resuspended in a stain solution to produce the antigen suspension (OIE Terrestrial Manual 2012). 16 1.9.3 Complement fixation test The complement fixation (CF) test has been used extensively for many years; however, it shows variable sensitivity (ranging from 20 to 60%), possibly reflecting differences in techniques for antigen production, and poor reproducibility . In addition , it has been demonstrated that the CF assay fa ils to detect a significant proportion of carrier cattle (Bradway et al., 2001 ). It is also uncertain as to whether or not the CF test can identify antibodies in acutely infected animals prior to other assays (Molloy et al. , 1999; Coetzee et al., 2007). Therefore, the CF test is no longer recommended as a reliable assay for detecting infected animals (OIE Terrestria l Manual, 2012). The use of the CF test could result in infected cattle being assigned a negative result, which could lead to introduction of persistently infected animals with false-negative resu lts into populations of completely na"lve cattle (Coetzee et al., 2007). 1.9.4 Indirect fluorescent antibody test Other serological tests are generally preferred to the IFA test , because of the number of tests that can be performed daily by one operator. A serious problem encountered with the IFA test is non-specific fluorescence attributed to antibodies adhering to infected erythrocytes. Non-specific fluorescence due to antibodies adhering to infected erythrocytes can be reduced by washing the erythrocytes in an acidic glycine buffer before antigen smears are prepared (OIE Terrestrial Manual 2012). 1.9.5 Polymerase chain reaction Polymerase chain reaction (PCR) has been the most commonly used method for the diagnosis of A. marginale infections (Lew et al., 2002, Shkap et al., 2002). Most of the PCR assays target the msp4 and msp1a genes for differentiating strains of A. marginale, which is a useful method for tracking the origin of the outbreak (Bowie et al., 2002; Lew et al. , 2002, de la Fuente et al. , 2005d; 2007a; Mtshali et al., 2007, Cabezas- Cruz et al., 2013; Mutshembele et al., 2014). 17 Polymerase chain reaction has been shown to reliably detect Anaplasma at the lowest levels of persistent rickettsemia (Lew et al., 2002, Shkap et al., 2002). However, other studies have shown that PCR fails to detect the infection of A. marginale at low levels or during the early stages of infection (Malad et al. , 2009). The molecular diagnosis of Anaplasma infection by PCR using whole blood has become readily available. Various genes are useful for organism identification at genus and species level. Among the most used are the phylogenetically reliable marker genes such as 16S RNA (Warner and Dawson, 1996; Wen et al., 2002), gltA (lnokuma et al., 2001 ) and groEL (Yu et al., 2001 ). Polymerase chain reaction assays targeted at the Anaplasma msp4 and/or msp 1a genes have been used to differentiate isolates of A. marginale, which is useful to track the origin of an outbreak, and to differentiate between different species of Anaplasma such as A. marginale and A. centrale (de la Fuente et al., 2001 a; Lew et al. , 2002). Because msp5 expresses epitopes that are conserved among widely divergent strains of A. marginale, A. centrale and A. ovis (Visser et al., 1992) and even A. phagocytophilum (Alleman et al., 2006), a PCR assay targeted only at the msp5 gene would be expected to have significant specificity issues when used in populations where animals could be infected with other Anaplasma spp. Amplification of related organisms by non-specific primers has been shown to result in false-positive reactions. However, sequencing and sequences comparison may be a solution for this problem. Conversely, false-negatives may occur if extraction procedures fail to remove PCR inhibitors present in a blood sample or if the level of circulating rickettsemia falls below the level of assay. The main problem with the above genes is that they are conserved at species levels so strains identification becomes challenging or impossible using these genes. An alternative is the use of variable genes. The msp 1a gene for A. marginale have become the most used tool for testing strains variability in these two pathogens (Palmer et al. , 2001 ; Almazan et al., 2008; Ruybal eta/. , 2009). 18 Nucleic-acid-based tests to detect A. marginale infection in carrier cattle have been developed although not yet fu lly val idated. The analytical sensitivity of PCR-based methods has been estimated at 0.0001 % infected erythrocytes, but at this level only a proportion of carrier cattle would be detected. A nested PCR has been used to identify A. marginale in carrier cattle with a capability of identifying as few as 30 infected erythrocytes per ml of blood, well below the lowest levels in carriers. However, nested PCR poses significant quality control and specificity problems for routine use (Torioni De Echaide et al. , 1998). Real-time PCR based on msp1b gene was successfully developed for detection and quantification of A. marginale DNA in blood of naturally infected cattle (Carelli et al. , 2007; Decaro et al., 2008; Reinbold et al. , 2010). The assay was shown to be distinctively sensitive and specific as there were no cross-reactions with other haemoparasites of ruminants (A. centrale, A. bovis, A. phagocytophilum, B. bigemina and T. buffelt). Then A. marginale real-time PCR was modified by adding a primer/probe set specific for A. centrale, thus obtaining a duplex assay for simultaneous detection of both Anaplasma spp. in the same reaction. Duplex real-time PCR has proven to be a powerful tool for differentiating between closely related infectious agents (Decaro et al. , 2006a; 2006b) and between vaccine and field strains of the same genotype (Decaro et al., 2006c; Elia et al., 2008). 1.10 Treatment, prevention and control strategies Control measures for anaplasmosis vary with different geographic locations, including maintenance of anaplasmosis free herds, control of tick vectors, administration of antibiotics and vaccination (de Waal, 2000). Dairy and beef cattle farmers have relied on dipping measures for control of tick infestation; however, in areas where tick vectors are well-established, the continuous exposure to ticks leads to endemic stability situation (de Waal, 2000). Vector control is labour intensive, expensive and environmental pollution is also a major concern. 19 Prophylaxis has been through administration of antibiotics. Chemotherapy is intended for prevention of clinical anaplasmosis but it does not prevent cattle from becoming persistently infected with A. marginale; however, cattle receiving antibiotics therapy may not be cleared of infection. Tetracycline administration is accompanied by disadvantages of expenses, and the demand of continuous feed ing and also the risk of development of resistant Anaplasma organisms, although the resistance of A. marginale to antibiotics has not been reported (de Waal, 2000). Kocan et al. , (2003) reviewed the progression and history of vaccine development which included live and killed vaccines. Both vaccines have relied on A. marginale infected bovine erythrocytes as source of antigen. The live and killed vaccines induce protective immunity that diminishes the cl inical signs, but does not prevent cattle from becoming persistently infected with A. marginale. Live vaccine (A centrale), which was introduced by Sir Arnold Theiler has been used widely in South Africa but the vaccination renders partial protection against A. marginale and its success vary according to genotypes of A. marginale (Brown et al. , 1998; de Waal, 2000). In contrast, trials conducted in South America and Africa with heterologous strains showed low efficacy of this vaccine (Turton et al., 1998; Bock and de Vos, 2001 ; Dark et al. , 2011 ). Major surface proteins (MSP) have been manipulatively used experimentally as vaccine candidates against A. marginale infections (Palmer et al., 1987; Allred et al. , 1990). MSP1 a has an ability to induce T-cell response and contains conserved B-cell epitope in the repeated peptides that is recognized by immunized and protected cattle (Kocan et al. , 2003, de la Fuente et al. , 2003b ). Experiments with recombinant MSP1 a have shown partial protection against anaplasmosis in cattle (Torina et al., 2014). 20 Recent studies have identified members of the pfam01617 from msp2 family to be possible vaccine candidates because most of them are found in cross-linked surface antigen complexes with the other members of pfam01617 encoding conserved outer membrane proteins, which are expressed in A. marginale (Noh et al. , 2006; Dark et al., 2011 ). 1.11 Anap/asma marginale major surface proteins and their role in host-vector- pathogen interactions The outer membrane proteins of A. marginale have been the focus of direct research to obtain an improved vaccine against bovine anaplasmosis (Palmer et al. , 1999). Immunization with purified outer membranes induces protection against acute A. marginale infection and disease (Tebele et al. , 1991 ). Various major outer membranes have been described based on the proteomic and genomic approach , and 21 proteins were identified within the outer membrane immunogen (Lopez et al. , 2005). Some well- characterized outer membranes proteins designated major surface proteins (MSPs): msp1a, msp1b, msp2, msp3, msp4 and msp5 were evaluated as potential candidates for antigens of vaccine production and diagnostic evaluations (Visser et al. , 1992; Oberle et al. , 1993; McGarey et al., 1994; Alleman and Barbet, 1996; Kano et al., 2002; Garcia-Garcia et al. , 2004a). 21 1.11.1 The MSP1 Complex The relationship of the rickettsia with the bovine host is complex as it has interaction with the erythrocyte and endothelial cells (Carreno et al., 2007). A number of rickettsial protein and the bovine immune system have not been studied in detai l yet, there is no final solution (vaccine or otherwise). Several major surface proteins (MSPs) have been identified in Anaplasma spp., which have been most extensively characterized in A. marginale (Palmer et al. , 1985; de la Fuente et al. , 2001 a, 2005b; Kocan et al. , 2003; 2004; 2008). Anaplasma MSPs are involved in interactions with both vertebrates and invertebrates host cells (de la Fuente et al. , 2001a; 2005a; Kocan et al., 2003; 2004; Brayton et al. , 2005; Dunning Hotopp et al., 2006; Nelson et al., 2008), and are likely to evolve more rapidly than other genes because they are subjected to selective pressures exerted by host immune systems. Six MPSs have been identified in A. marginale from cattle and ticks of which three, msp1a, msp4 and msp5, are coded by single gene and do not vary within isolates (Bowie et al. , 2002). The other three, msp1b, msp2 and msp3, are from multigene families and may vary antigenetically in persistently infected cattle (Barbet et al. , 2000; Kocan et al. , 2000; Bowie et al., 2002). Major surface protein 1 is a heterodimer composed of two structurally unrelated polypeptides: msp1a, (100kDa), which is encoded by a single gene, msp1a (Lew et al. , 2002) and msp1b (105 kDa), which is encoded by at least two genes, msp1f31 and msp1f32 (Barbet et al., 1987; Viseshakul et al. , 2000; Camacho-Nuez et al., 2000; Bowie et al., 2002; Macmillan et al. , 2006). The msp1a, and msp1b proteins form non-covalent dimers and are exposed on the surface of A. marginale (Barbet et al., 1987). However, only a single msp1b protein, msp1b1 , was identified within the MSP1 complex (Macmillan et al. , 2006). 22 The molecular weight of msp1a varies in size among isolates due to different numbers of tandemly repeated 23-31 amino acid peptides and has been used for identification of geographic strains (Allred et al. , 1990; de la Fuente et al. , 2001 a; 2003a; 2005a; 2007a). The msp1a, tandem repeats are located after a conserved decapeptide in the amino terminal region of the protein and are exposed extracellularly for interaction with host cell receptors (de la Fuente et al., 2003a). The frequency of variable amino acid positions within geographic isolates is higher in this region than in the rest of the protein (de la Fuente et al. , 2001 a). Functionally, msp1a was shown to be an adhesin for bovine erythrocytes and tick cells and the adhesin domain was identified on the variable N-terminal region contain ing the repeated peptides (McGarey and Allred, 1994; McGarey et al. , 1994; de la Fuente et al., 2001 b). It was also shown to be involved in the transmission of A. marginale by Dermacentor spp. (de la Fuente et al. , 2001 c) and to be differentially regulated in tick cells and bovine erythrocytes (Garcia-Garcia et al., 2004b). The msp1a, although variable in the number of repeated peptides, induces strong T-cell responses and contains a conserved B-cell epitope in the repeated peptides that is recognized by immunized and protected cattle (Kocan et al., 2003; de la Fuente et al., 2005a). The msp1a also contains Th1 cell epitopes in the conserved region, which may be involved in immunoprotection (Brown et al., 2001). The msp1b, encoded by at least two genes, msp1b and msp1b2, is polymorphic among geographic isolates of A. marginale (Barbet et al. , 1987; Camacho Nuez et al. , 2000; Viseshakul et al. , 2000; Bowie et al., 2002). Although msp1b is encoded by a multigene family, only small variations in protein sequences of msp1b and msp1b2 were observed during the life cycle of the Rickettsia in cattle and ticks (Bowie et al. , 2002). This protein, which forms a complex with msp1a, is an adhesin for bovine erythrocytes (McGarey and Allred , 1994; McGarey et al., 1994). However, MSP1b was recently demonstrated to be adhesin only for bovine erythrocytes and did not prove to be an adhesin for tick cells (de la Fuente et al. , 2001 a). 23 The A. marginale surface is characterized by the presence of two highly abundant and closely related outer membrane proteins Major Surface Protein 2 (msp2) and (msp3) (Brayton et al., 2001 ). The predominant immune responses are generated against these two proteins (McGuire et al. , 1991 ; Brown et al. , 2001 ; 2003). However, both msp2 and msp3 are highly antigenically variable, both within an infection and between strains (McGuire et al., 1984; French et al., 1998; 1999; Rodriguez et al., 2005). Thus, while antibody response to msp2 and msp3 antigenic variants plays a key role in how persistent infection is established and the population strain structure, these abundant surface proteins are not targets for development of a widely cross-protective vaccine and anti- msp2/msp3 immune responses do not associate with protective efficacy of the outer membrane vaccine (Palmer et al. , 2009; Noh et al. , 2010). Using genomic and proteomic approaches, the minor components of the outer membrane protein immunogen have been identified (Noh et al., 2006; 2008; Brayton et al. , 2005; Lopez et al. , 2005; 2008; Agnes et al. , 2011 ). Although markedly less abundant, these minor proteins are invariant during infection and highly conserved among strains-thus representing much more attractive targets fo r vaccine development. Importantly, the proteomic identification within the outer membrane immunogen and the bioinformatics prediction of surface localization was confirmed for a subset of these proteins by surface-specific cross-linking (Noh et al. , 2008). The isolated cross-linked surface protein complex induced protection equal to that of the native outer membrane immunogen (Noh et al., 2008). 24 The msp2 and msp3 are both encoded by large polymorphic, multigene families (Palmer et al. , 1994; Alleman et al., 1997). The msp2 sequence and antigenic composition varies during cyclic rickettsemia in cattle (French et al. , 1998; 1999; Barbet et al., 2001) and in persistently infected ticks (de la Fuente et al. , 2001 a). MSP2 is encoded on a polycistronic mRNA. The msp2 gene within the expression site is polymorphic. The msp2 encodes numerous amino acid sequence variants selected in bovine erythrocytic and tick salivary gland populations of A. marginale (French et al. , 1998; 1999; Barbet et al., 2000; Brayton et al., 2001 ; de la Fuente et al. , 2001 a; Meeus and Barbet, 2001 ). The msp3 also varies in antigenic properties and structure among geographic isolates (Alleman and Barbet, 1996). The msp2 and msp3 are involved in the induction of a protective bovine immune response to A. marginale (Palmer et al. , 1999). The msp4 and msp5 are encoded by single copy genes. Although msp4 is highly conserved (Oberle and Barbet, 1993; de al Fuente et al. , 2003a), information about its function is not available. The msp5 is highly conserved surface protein that has been proven effective as a diagnostic antigen and used in a competitive enzyme linked immunosorbent assay (cELISA) commercially available in the United States (Torioni De Echaide et al. , 1998). The function of msp5 is also unknown. The msp2 operon-associated genes OpAG1 , OpAG2, and OpAG3, have been identified in A. marginale and may encode for surface proteins (Lohr et al. , 2002). 25 1.11.2 Vector-pathogen relationships 1.11.2.1 Anaplasma MSPs and vector -pathogen interactions The evolutionary history of vector-pathogen interactions could be reflected in the sequence variation of the Anaplasma major surface proteins. Previous studies demonstrated that A. marginale msp1a, but not msp4 is under positive selection pressure (de la Fuente et al., 2003a). Initial analysis of A. marginale msp1a and Dermacentor variabilis 16S rDNA sequences from various areas of the United States suggested tick-pathogen co-evolution (de la Fuente et al., 2001 d) consistent with the biological function of MSP1a involved in the transmission of A. marginale by ticks (de la Fuente et al. , 2001 c). However, the genetic diversity of sequences complicates the study of tick-pathogen co-evolution (de la Fuente et al. , 2010). Anaplasma marginale strains are apparently not transmissible by ticks and rely on mechanical transmission for completion of the life cycle in nature (de la Fuente et al., 2005d; Hornok et al., 2008). These facts pose the question of what evolutionary adaptations may have occurred to insure an efficient mechanical transmission of non- tick-transmissible A. marginale strains. Recently, Scoles et al. , (2005) concluded that biological transmission by ticks is more efficient than mechanical transmission of A. marginale (de la Fuente et al. , 2010). The adhesin domain of A. marginale msp1a has been identified on the extracellular N- terminal region of the protein that conta ins the repeated peptides (de la Fuente et al., 2003b). The binding of msp1a to tick cell extract (TCE) was observed within the msp1a tandem repeats, the negatively charged amino acids, aspartic acid (0) and glutamic acid (E), at position 20 demonstrated that peptide containing acidic amino acids D or E at position 20 bound to TCE, while peptides with G as the 201h amino acid were not adhesive to TCE (de la Fuente et al., 2003b). 26 1.12 Phylogenetic relationships of geographic isolates of A. marginale A phylogenetic tree is an estimate of the relationships among taxa (or sequences) and their hypothetical common ancestors (Nei and Kumar 2000; Felsenstein 2004; Hall 2011 ). Today most phylogenetic trees are built from molecular data: DNA or protein sequences. Originally, the purpose of most molecular phylogenetic trees was to estimate the relationships among the species represented by those sequences, but today the purposes have expanded to include understanding the relationships among the sequences themselves without regard to the host species, inferring the functions of genes that have not been studied experimentally (Hall et al., 2009), and elucidating mechanisms that lead to microbial outbreaks (Hall and Barlow 2006) among many others. Building a phylogenetic tree requires four distinct steps : (Step 1) identify and acquire a set of homologous DNA or protein sequences, (Step 2) align those sequences, (Step 3) estimate a tree from the aligned sequences, and (Step 4) present that tree in such a way as to clearly convey the relevant information to others (Hall, 2013). Phylogenetic analysis of A. marginale geographic isolates from the United States was performed using the genes msp1a and msp4 (de la Fuente et al., 2001d). The results of these analyses strongly support a southeastern clade of A. marginale composed of isolates from Virginia and Florida. Analysis of 16S rDNA fragment sequences from the tick vector of A. marginale, D. variabilis, from various areas of the United States was performed and suggested coevolution of the vector and pathogen (de la Fuente et al., 2001d). Isolates of A. marginale from the United States also grouped into two clades, southern clade consisting of isolates from Florida, Mississippi , and Virgin ia, and a west-central clade consisting of isolates from California, Idaho, Illinois, Oklahoma, and Texas. Although little phylogeographic resolution was detected within any of these higher clades, msp4 sequences appear to be a good genetic marker for inferring phylogeographic patterns of isolates of A. marginale on a broad geographic scale . 27 In contrast to the phylogeographic resolution provided by msp4, DNA and protein sequence variation from msp1a representing 20 New World isolates of A. marginale failed to provide phylogeographic resolution (de la Fuente et al. , 2007b). Most variation in msp1a sequences appeared unique to a given isolate. In fact, simi lar DNA sequence variation in msp1a was detected within isolates from Idaho and Florida and from Idaho and Argentina. These results suggest that the msp 1a sequence may be rapidly evolving and that the msp1a gene may provide phylogeographic information only when numerous msp1a sequences from a given area are included in the analysis (de la Fuente et al. , 2007a). The analysis of msp 1a DNA and protein sequences demonstrated extensive genotypic variation among Oklahoma isolates of A. marginale and failed to provide phylogeographic resolution within Oklahoma or on a broader scale, including isolates from other United States and Latin America. Furthermore, analysis of codon and amino acid changes over the msp1a and msp4 phylogenies provided evidence that msp1a but not msp4, is under positive selection pressure. These results suggest that even if msp1a sequences are rapidly evolving, MSP1a genotypes reflect the history of cattle movement more than the geographic distribution of A. marginale isolates. Research analysis suggest that different A. marginale genotypes are maintained within a herd in an area of endemic infection by independent transmission events and that infection with more than one genotype per host is prevented, a phenomenon described as infection exclusion (de la Fuente et al., 2003a). Therefore, if cattle movements import a new A. marginale genotype, it could be established by mechanical and/ or biological transmission to susceptible cattle. In regions with few cattle introductions, like Australia, little genotypic variation are found within A. marginale isolates (de la Fuente et al. , 2005a). 28 CHAPTER 2 OBJECTIVES OF THE STUDY 2.1 Statement of the problem Anaplasma marginale genotypes have been well characterized and are highly variable in endemic areas worldwide (de la Fuente et al., 2010) . Anaplasma marginale is quite host specific for ruminants and anaplasmosis occurs primarily in cattle; other ruminants may serve as reservoirs of infection. While A. marginale is transm itted biologically by ticks, mechanical transmission by blood-contaminated mouthparts of biting flies or fomites also frequently occurs. Mechanical transmission may be the only means of spreading anaplasmosis in areas where tick vectors are absent or are unable to transmit the local A. marginale strain (Kocan et al., 2010). Immunization of cattle with affinity- purified native MSP1 complex induced protective immunity in cattle that received homologous or heterologous challenge with A. marginale geographic isolates (Palmer et al., 1987; 1989). The msp1a has been shown to have a neutralization sensitive epitope (Palmer et al., 1987) and to be an A. marginale adhesin for both bovine erythrocytes and tick cells, while msp1b1 is an adhesin only for bovine erythrocytes (McGarey et al., 1994; McGarey and Allred , 1994; de la Fuente et al. , 2010). Killed vaccine marketed previously in the USA used A. marginale antigen that was partially purified from bovine erythrocytes (Brock et al. , 1965; Hart et al., 1981 ; Mccorkle-Shirley et al., 1985). This vaccine was used effectively until it was removed from the market in 1999 (Kocan et al., 2010). These blood-derived killed vaccines reduced clinical anaplasmosis and were expensive to produce, difficult to standardize and often not cross-protective in widely separated geographic areas with different endemic A. marginale isolates. The blood-derived vaccines also bore risk of being contaminated with bovine cells or pathogens that frequently cause persistent infections in cattle (Palmer, 1989; Kocan et al., 2000). 29 Despite the importance of anaplasmosis in South Africa and other African countries, A. marginale strains have not been genetically characterized in Africa (Mtshali et al. , 2007). Vaccination experiments with recombinant msp1a have resulted in partial protection against clinical anaplasmosis in cattle and reduced infection levels in ticks, thus supporting the inclusion of msp1 a in vaccines for the control of bovine anaplasmosis (Kocan et al., 2010). The phylogenetic analysis of A. marginale strains using MSPs has been recently reviewed. These analyses suggest that MSPs may not be good markers for biogeographically studies on a global scale. However, they may be useful for strain comparison in given regions and could provide information about the evolution of host- pathogen and vector-pathogen relationships (Cabezas-Cruz et al., 2013). While a universal vaccine has yet to be developed for anaplasmosis, preliminary analysis of the phylogeographic relationships of A. marginale based on two major surface proteins (msp1a and msp4) has shown phylogeographic partitioning of parasite isolates in the United States (de la Fuente et al. , 2007a). This information supports the inclusion of several geographic isolates of A. marginale in development of vaccine formulations for the United States (Kocan et al., 2010). Recent research has focused on MSPs that may be used to elucidate phylogeographic patterns of A. marginale (de la Fuente et al. , 2001 a) and that are involved in interactions with vertebrate and invertebrate host cells (McGarey et al., 1994; McGarey and Allred, 1994; de la Fuente et al. , 2001a). 30 These MSPs involved in host-pathogen interactions may evolve more rapidly than other nuclear genes because of selective pressures exerted by host immune systems (de la Fuente et al., 2010). Of the six A. marginale MSPs that have been identified and characterized , only three (msp1a, msp4 and msp5) are each encoded by a single gene. Because these MSPs do not appear to undergo antigenic variation in cattle or ticks, they were reported to be more stable genes for phylogenetic studies (Bowie et al. , 2002). Of these three MSPs, msp1a has been reported to be an adhesin for bovine erythrocytes and tick cells and to effect infection and transmission of A. marginale by Dermacentor spp. ticks (McGarey et al. , 1994; McGarey and Allred , 1994; de la Fuente et al., 2001 b, 2001 c). Although the specific function of msp4 is currently is not known , the previous analysis of the gene from A. marginale isolates detected sufficient sequence variation to support its use in phylogeographic studies (de la Fuente et al., 2001 a). Therefore, msp4 warranted further testing for use in detection of phylogeographic patterns among A. marginale isolates. In a previous report of the phylogeographic relationships of A. marginale isolates from the United States (de la Fuente et al., 2001a). It was found that DNA sequence variation in msp4 provided phylogenetic resolution for intraspecific relationships among isolates (de la Fuente et al. , 2002b). In addition, when DNA sequence variation of msp1a was interpreted in combination with msp4 phylogeny, important phylogeographic information became apparent. This study utilised MSPs to explore the epidemiology and phylogeny of A. marginale causing anaplasmosis in selected South African provinces. 31 2.2Aim of the study The aim of this study is to determine the prevalence, genetic distribution and phylogenetic relationship of A. marginale infecting South African-cattle by using msp1a and msp4 gene 2.3 Objectives 2.3.1 To determine the prevalence of A. marginale infecting cattle in South Africa by using msp1a as a target gene. 2.3.2 To establish genetic diversity of A. marginale infecting cattle in South Africa using msp1a as marker gene. 2.3.3 To determine phylogenetic relationship of A. marginale infecting South African cattle by using msp1a and msp4 genes. 32 CHAPTER 3 MOLECULAR DETECTION OF ANAPLASMA MARG/NALE INFECTING CATTLE IN SOUTH AFRICA BY PCR TARGETING msp1a GENE 3.1 Introduction Bovine anaplasmosis is a tick-borne hemoparasitic disease caused by Anaplasma marginale and is the most widely distributed disease of cattle in South Africa (de Waal, 2000; Ndou et al., 2010). Economic losses associated with anaplasmosis in cattle include impaired production , mortality and expenses in control measures (Regassa et al., 2003). Ticks and the diseases they transmit have been identified as the major cause of widespread morbidity and mortality in cattle kept by smallfarm holders in the semiarid areas of South Africa (Dold and Cocks, 2001 ; Mapiye et al. , 2009). Poor cattle health management, resistance of ticks to most acaricides and the use of inappropriate cattle breeds (Dold and Cocks, 2001 ; Marufu et al. , 2011 ) have increased the prevalence of ticks and tick-borne diseases (TBD) in smallholder cattle herds. Polymerase chain reaction detection methods have been developed. These tests/assays are extremely sensitive and specific in the detection of A. marginale infections in cattle (Lew et al., 2002; de la Fuente et al., 2005b; Malad et al. , 2006). Some of these assays, such as the real-time PCR, were developed to enable simultaneous detection and quantification of the A. marginale DNA in bovine blood. This is essential in supporting the clinical diagnosis, assessing the carrier status of the cattle and evaluating the efficacy of vaccines and antirickettsial drugs (Carelli et al., 2007). 33 Among the genes encoding for major surface proteins in A. marginale, msp1a has been extensively used for strain characterization (Cabezas-Cruz et al. , 2013). The protein msp 1a is involved in the interaction of the bacterium with vertebrate and invertebrate host cells (de la Fuente et al. , 2010). Several strains of A. marginale have been identified worldwide and these strains differ in their morphology, msp1a amino acid sequence, antigenic characteristics, and ability to be transmitted by ticks (de la Fuente et al., 2007a; Estrada-Pena et al., 2009). The msp 1a gene is thus a good candidate for the development of a generic A. marginale genotyping assay that could be applied globally and also to track strains (Lew et al. , 2002). The objective of the study was to determine the prevalence of A. marginale infecting cattle from Limpopo, Mpumalanga, North West, Gauteng, KwaZulu-Natal, Eastern Cape, Western Cape and Northern Cape provinces of South Africa by PCR using the msp1a gene. 3.2 Materials and methods 3.2.1. Study site and sample collection Two hundred and eighty samples were collected from cattle in several South African provinces and the exact locations with coordinates are provided in Figure 5 and Table 4. Farms with different productive systems were chosen. Cattle were selected on a random basis following their owner's approval for the collection of blood samples. Blood samples were obtained from cattle and no stratification according to age or sex was applied. No vaccination histories were available for these animals. Tail venupuncture was performed using an 18 gauge needle and blood samples were collected into sterile vacutainer tubes (10 ml) with EDTA and were kept at 4 °C until delivered at the laboratory. 34 3.2.2 DNA extraction The genomic DNA was extracted from cattle blood samples using ZR Genomic DNA rM Tissue Miniprep (Zymo Research, CA, USA). Total volume of blood sample was adjusted in a microcentrifuge tube with water to 100 µI prior to adding 95 µI of 2x Digestion buffer and 5 µI of Proteinase K. The mixture was thoroughly mixed and then incubated at 55 °C for 20 minutes. A volume of 700 µI of genomic lysis buffer was added to the tube and mixed thoroughly by vortexing. The mixture was carefully transferred to a Zymo-Spin™ llC column in a collection tube and centrifuged at 10,000 x g for one minute. A volume of 200 µI of DNA pre-wash buffer was added to the spin column in a new collection tube and centrifuged at 10,000 x g for one minute. A volume of 400 µI of g-DNA wash buffer was added to the spin column and centrifuged at 10, 000 x g for one minute. The spin columns were carefully transferred to a clean 1.5 ml microcentrifuge tube and 100 µI of DNA Elution buffer added to the spin column. The microcentrifuge tube was incubated for 5 minutes at room temperature, and then centrifuged at 8000 rpm for 30 seconds to elute the DNA. The eluted DNA was stored at -20 °C until used for studies. The concentration of DNA was measured using a NanoDrop® ND-1000 (NanoDrop Technologies Inc., Wi lmington, USA). 3.2.3. Polymerase chain reaction (PCR) A specific set of primers was used to amplify msp 1a gene. A PCR was performed using the primers that were previously designed to amplify a fragment of msp1a gene, contain ing the tandem repeat region. Primers were synthesized and supplied by lnqaba Biotech (lnqaba Biotechnical Industries (Pty) Ltd, Pretoria, Gauteng, South Africa). The PCR reagents mixture consisted of total volume of 25 µI of Thermo Scientific DreamTaq Green PCR Master Mix (2X), 10 µM of forward and reverse primers, 1 µg Template DNA in 0.2 ml PCR tubes. 35 The PCR mixture was then amplified by DNA thermocycler (Bio-Rad T100™ DNA Thermal Cycler, Bio-Rad Laboratories, Johannesburg, South Africa). The amplification process involved an initial denaturation at 94 °C for 3 minutes followed by 40 cycles of denaturation at 94 °C for 1 minute at 94 °C, annealing at 65 °C for 1 minute , and an extension at 72 °C for 2 minutes. A cycle of a final extension was done at 72 °C for 7 minutes. The amplified products were then separated by electrophoresis using 1 % agarose gel immersed in TSE buffer (89 mM Tris-Borate, 2 mM EDTA, pH 8). One kilo- base pair ladder marker was used to determine the size of the PCR products after staining with 0.5 µg/ml GelRed (Life Sciences, Fermentas GmbH, Germany) and visualization under UV illumination. Gel documentation was done and taking a photograph using a camera (Figure 3). Table 3: Oligonucleotide sequences of primers used in this study Primer designation Oligonucleotide (5'-3') References 1733F TGTGCTTATGGCAGACATTTCC Lew et al. , 2002 2957R AAACCTTGTAGCCCCAACTTATCC Lew et al., 2002 3.2.4. Statistical analysis Statistical analysis of the results were performed at the Department of Biostatistics and Statistics at the Medical Research Council , Pretoria Unit using the Fisher's exact and Chi-square test to test whether the prevalence is associated with the type of farm while chi-squared test was used to evaluate any significant difference (P=0.0001 ). The confidence intervals and standard errors of the prevalence of Anaplasma marginale infections in cattle were calculated at 95% using Kruskal-Wallis equality of population rank test. Pearson chi-square test was used to analyse PCR results based on frequency, row, column and cell percentage (Appendix 1) . 36 3.3. Results 3.3.1 Detection of A. marginale infections Two hundred and eighty blood samples were analysed by PCR technique. The primers 1733F and 2957R were used to amplify a fragment of msp 1a gene, contain ing the tandem repeat region which varies among isolates. In this study PCR amplicons of variable sizes were obtained, ranging from 630 to 1100 bp corresponding to msp1a gene of different molecular weight as shown previously by Lew (2002). Two hundred and nine cattle blood samples out of 280 were PCR positive for A. marginale infection as shown in Appendix 1. 3.3.2 Molecular prevalence of A. margina/e isolates infecting cattle in South African provinces The PCR results of A. marginale infection ranged from 0% to 100% for samples from Mpumalanga and Gauteng. Eastern Cape comprised of 100%, KwaZulu-Natal 87.27%, Western Cape 84.62%, Limpopo 80% and North West 78. 79% positive infectious rates. None of the Northern Cape samples were positive by PCR (Appendix 1 ). 3.3.3 Statistical prevalence of A. marginale isolates infecting cattle in South African provinces Statistical analysis was conducted to obtain the prevalence rates using the PCR results. The highest relative prevalence was observed in KwaZulu-Natal with 23.0%, second Eastern Cape 19.1% , then Limpopo 15.3%, North West 12.4%, Western Cape 10.5%, Mpumalanga 10.1% and Gauteng 9.6%. These were obtained by considering relative rates of prevalence with in the province. There was no significant statistical association between prevalence of A. marginale with the type of farm (communal or commercial). 37 PCR products with single bands as shown in Figure 3 were sequenced. The single PCR products from the seven different geographical regions including Limpopo, Mpumalanga, North West, Gauteng, KwaZulu-Natal, Eastern Cape and Western Cape of South African provinces A. marginale were sequenced and forty four gene sequences were submitted to GenBank (Appendix 2). BLAST analysis confirmed that the DNA results were indeed A. marginale msp1a gene and showed high identity (90-100%) with A. marginale msp1a deposited in GenBank. 38 Although A. marginale was detected in some of the samples as indicated by PCR results, most of the positive PCR samples from Mpumalanga, KwaZulu-Natal and Eastern Cape provinces, had non-specific bands and were not further analysed (Figure 3). M 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 10000 5000 3000 2000 1200 850 500 300 100 Figure 3: Gel image of 1 % agarose gel electrophoresis of PCR amplification products obtained from A. marginale isolates from Western Cape Province in South Africa using 1733F and 2973R primers (Lew et al., 2002). Lane Marker (M): GeneRuler™ I Kb DNA ladder, ready-to-use; Lane 1-5, 7-8, 13-14, 16-18 indicate positive results and lane 6, 9-12, 15 indicate negative results; lane 19 negative control. 39 60.0 50.0 40.0 VJ 2 Ill 30.0 0 23.0 .!!1 19.1 20.0 .91 15.3 12.4 .Isl:l 9.6 10.1 10.5 e> 10.0 Ill E 0.0 ~ 0.0 L I I I I I I 0 e ~~ ~ c.l~ ~ ~o ~e; ~~~ .:f.e<} ~e ~e Q) ~ ~o <.J <.J 0 I..~ ~~ c ~ (I) ._e; <> ~~ ~~ ,l' ~~ I..~ ._(:' ~e; (ij ~'b. . ~ ~o .:t-e" > ~.:J. ~~ ~o~ ~ Cl. • Positiw Figure 4: Prevalence of A. marginale by province in South Africa. 40 3.4 Discussion It has been reported that bovine anaplasmosis is one of the most important cause of cattle mortalities in low-input farming areas in South Africa (Mapiye et al. , 2009). Polymerase chain reaction detection methods have been developed, which are extremely sensitive and specific for the detection of A. marginale infections in cattle (de la Fuente et al., 2005b; Molad et al. , 2006). It is crucial to use sensitive and efficient molecular techniques to generate accurate information on the prevalence of A. marginale, which are crucial not only for developing appropriate control measures but for providing an understanding of host resistance in different cattle genotypes (Marufu, 2014). In this study, a sensitive and specific msp1a PCR (Lew et al. , 2002) was used to determine the presence of A. marginale in eight different provinces of South Africa (Appendix 1) . The results indicate the presence of A. marginale infection in these provinces with an overall prevalence of 76% except for Northern Cape in which the prevalence was 0%. This prevalence may be considered as high as in other endemic regions of the world. Recently, 70% prevalence of A. marginale in a cattle herd was reported in Brazil (Pohl et al. , 2013). Another study by de la Fuente et al., (2005a) showed a prevalence range from 25 to 100% for A. marginale infecting cattle herds in Italy. Results of this study confirm that bovine anaplasmosis is endemic and highly prevalent in South Africa . The absence of positive samples in Northern Cape Province may be explained by special climate conditions in th is area that limits the survival of tick vectors (Mtshali and Mtshali , 2013). However, the prevalence of bovine anaplasmosis at certa in areas is a complex process also involving bovine population immunity to A. marginale, tick control practices, tick transmission, grazing system as well as the genetic variability of A. marginale at the population level (Hamou et al., 2012). 41 Despite high prevalence of A. marginale in the studied area, the veterinary authority (State Veterinary officials) has not reported clinical disease in cattle, which is an indication of endemic stabil ity . Endemic stability is an epidemiological state, in which clinical disease is scarce despite high levels of infection in the population (Coleman et al. , 2001 ; Jonsson et al., 2012). This is the first report of A. marginale prevalence at the national level based on molecular data. The results confirm endemicity and high prevalence of A. marginale in South Africa. Further investigations should address the influence of such high prevalence in cattle in South Africa . This will help in the understanding of the epidemiology and the use of appropriate control and prevention measures to reduce the high prevalence in cattle in South Africa before experiencing an outbreak. 42 CHAPTER4 EPIDEMIOLOGY AND EVOLUTION OF THE GENETIC VARIABILITY OF ANAPLASMA MARG/NALE IN SOUTH AFRICA This work has been published as follows: A.M. Mutshembele, A. Cabezas-Cruz, M.S. Mtshali, O.M.M. Thekisoe, R.C. Galindo and J. de la Fuente. 2014. Epidemiology and evolution of the genetic variabi lity of Anaplasma marginale in South Africa. Ticks and Tick -borne diseases, 5 (6): 624-631 . 4.1 Introduction Bovine anaplasmosis is a non-contagious tick-borne disease caused by infection of cattle with A. marginale, an obligate intraerythrocytic bacterium classified in the family Anaplasmataceae, order Rickettsiales (Dumler et al. , 2001 ). This pathogen is transmitted biologically by ticks, mechanically by biting insects and blood-contaminated fomites and from cow to calf via transplacental transmission (Aubry and Geale, 2011 ). Five tick species have been shown experimentally to transmit A. marginale in South Africa, including Rhipicephalus microplus, R. decoloratus, R. evertsi evertsi, R. simus and Hyalomma marginatum rufipes (as reviewed by de Waal. , 2000). Acute disease in cattle is characterized by weight loss, fever, abortion, low milk production and in some cases death; the animals that recover from the disease become persistently infected and serve as reservoir of infection for mechanical transmission and biological transmission by ticks (Kocan et al., 2003). Anaplasmosis is widespread in South Africa and, as estimated by de Waal. (2000), 99% of the total cattle populat ion is at risk of acquiring A. marginale infection . 43 Currently, antimicrobial drugs are not available for the elimination of persistent infections in cattle. Although the World Organization for Animal Health proposed the use of enrofloxacin, imidocarb, and oxytetracycline for the elimination of persistent A. marginale infections in cattle, these antimicrobial drugs were not found to eliminate the persistent A. marginale infections (Coetzee et al., 2005; 2006). Vaccines have been used as an alternative method for control of anaplasmosis. A live vaccine using A. marginale sub sp centrale (A. centrale), a subspecies of relatively low pathogenicity, is available in South Africa, Australia, Israel and Latin America but this vaccine has proven to be only partially effective and reports of vaccination failure are not uncommon (Kocan et al. , 2003). While anaplasmosis still constitutes a problem for cattle production in South Africa, sales of vaccines for bovine anaplasmosis in South Africa have reduced from 800, 000 to 200, 000 doses over 22 years (1976-1998) (de Waal. , 2000) . MSP1 a has been shown to have potential use as a vaccine antigen because this protein contains both neutralization sensitive (Palmer et al. , 1987; Allred et al., 1990) and immunodominant epitopes (Garcia-Garcia et al., 2004b). Recently, use of MSP1 a for vaccine development for A. marginale has regained new attention. The N-terminus tandem repeated reg ion of this protein was used in immunization trials in cattle against A. marginale (Torina et al., 2014) and laboratory animal models (Santos et al., 2013; Silvestre et al., 2014 ), and demonstrated promising results . MSP1 a is one of six MSPs that have been described in A. marginale. This protein is encoded by a single-copy gene, msp1a, which is conserved during the multiplication of the parasite in cattle and ticks (Kocan et al. , 2003) and has been useful for epidemiological studies of A. marginale in various regions of the world (Ruybal et al. , 2009; Almazan et al., 2008). 44 r-------------------------------------------------- MSP1 a contains tandem repeats at the N-terminus of the protein which present functional residues that serve as adhesins for bovine erythrocytes and tick cells, a prerequisite for infection of host cells (McGarey et al. , 1994; de la Fuente et al., 2003a). While the tandem repeats of MSP1a are highly variable, repeats are commonly represented among worldwide strains (Cabezas-Cruz et al., 201 3) and have been shown to evolve under positive selection (de la Fuente et al., 2003b). However, the specific codon positions that evolve under positive or negative selection have not been reported. For development of MSP1a-based vaccines, some characteristics of MSP1 a should be taken into consideration , including (i) the extant genetic variability of MSP1 a, (ii) the evolution of MSP1 a genetic diversity, and (iii) the conservative nature of some tandem repeats among different isolates. In this study molecular evidence is provided regarding the prevalence of A. marginale in eight of the nine South African provinces. Additionally, evolution of the genetic diversity of A. marginale msp1a in South Africa was studied, demonstrating that different codon positions of this gene evolved under positive or negative selection, likely due to immune selection and transmission fitness. Finally, these studies demonstrated the low variability of some tandem repeats commonly found among A. marginale strains. Collectively, results of this research will contribute toward development of new and novel vaccines for control of bovine anaplasmosis in South Africa and other regions of the world. 45 4.2 Materials and methods 4.2.1 Study site and sample collection Blood samples were collected for these studies from May 2011 to July 2013 in 26 districts and municipalities from eight South African provinces (Figure 5). Coordinates of the collection sites are provided in Table 4, as well as the farm production systems. Cattle for these studies were randomly selected and blood samples were collected only from adult animals after the owner's consent. While information regarding the age or sex of the animals was not recorded, vaccination histories were not available for these animals. Blood samples were collected by tail venupuncture using an 18 gauge needle and sterile 10 ml vacutainer EDTA tubes and stored at 4 °C. 46 - A. marginale epidemic areas - A. marginale endemic areas D Areas free of A. marginale ~ Lesotho, area not included in the study Ticks species involved in A. margina/e transmission: + (R. evertsi everts1) * (R. microplus and R. decolaratus ) Figure 5: Map of South African areas included in the study. Map of South Africa showing the provinces that were included in the study (Limpopo, Mpumalanga, North West, Gauteng, KwaZulu- Natal, Eastern Cape, Western Cape, and Northern Cape). Endemic, epidemic, and A. margina/e-free areas are coloured differentially (data collected from de Waal, 2000). The main tick species involved in the transmission of A. marginale in the sampled areas are shown: Rhipicephalus microplus, R. decoloratus, and R. evertsi evertsi (data collected from de Waal, 2000). 47 Table 4: Sampling sites and their coordinates in South African provinces. South African provinces Sampling sites Map coordinates Limpopo Makgodu, Masehlong, Chloen 23° 36'S, 29° 18'E, farm, Aganang Municipality 23° 32'S, 29° 03'E, 23° 45'S, 28° 51 'E, Mpumalanga Ehlanzeni South District 25° 27'S, 30° 58'59"E Gauteng West Rand District, Merafong 26° 20' 1" S, 27° 19'39"E Municipality, Khutsong South and 26° 22'S, 27° 24'E Carltonville North West Moretele district, Maubane 25° 16'S, 28° 15'E KwaZulu-Natal Pietermaritzburg Chota, Umhlati 29°29'17.16"S, District, Albert Falls, Shallow drift, 30°26'27. 78"E, UMgungundlovu District, 29°28' 33.5" S, Richmond Municipality, Ndaleni 30°26' 11 .3" E Dip Tank 29° 52' 55.3" S, 30° 40' 56.2" Eastern Cape Amathole District, Nkokobe 26°38'39"E, Municipality (Fort Beaufort, Alice) 32°46'52"S, and Middledrift and Nxuba 26°51 '25"E, 32'41 .12S, Municipality (Bedford, Adelaide) 27°11 '58"E, 32° 52 '37"S, 26°26'40"E, 32°44'25"S, 26° 18'18"E, 32° 42'04"S Western Cape Stellenbosch district, Boland 33°44'28.6"S, 18°59' 3"E Northern Cape John Taolo Gaetsewe District, 27°18'53"S, Ga-Segonyana Municipality, 23° 42 '15"E Kuruman, Zero Farm 48 4.2.2 DNA extraction Genomic DNA was extracted from cattle blood samples using ZR Genomic DNA r M Tissue Miniprep (Zymo Research, CA, USA). DNA was resuspended in DNA elution buffer and stored at -20 °C. The concentration of DNA was determined using the NanoDrop® ND-1000 (NanoDrop Technologies Inc., Wi lmington, USA). 4.2.3 A. marginale species-specific PCR Specific set of primers were used to ampl ify msp 1a 1733F- (5'- TGTGCTTATGGCAGACATTTCC-3'), 2957R- (5'- AAACCTTGTAGCCCCAACTTATCC- 3') - gene (Lew et al. , 2002). Primers were synthesized and supplied by lnqaba Biotech (lnqaba Biotechnical Industries (Pty) Ltd , Pretoria, Gauteng, South Africa). PCR reactions were prepared using Master Mix (Thermo Scientific DreamTaq Green PCR) , 0.1-1 .0 µM of Forward and Reverse primers and 1 µg of DNA template in 25 µI final volume. The amplification cycles consisted of 40 cycles of 1 minute at 94 °C, 1 minute at 65 °C, and 2 minutes at 72 °C. Amplified products were separated in 1% TBE (89 mM Tris-Borate, 2 mM EDTA, pH 8) agarose gel using 1 Kb ladder as a DNA size marker (1 Kb DNA ladder, Life Sciences, Fermentas GmbH, Germany). DNA was visualized by gel staining in 0.5 µg/ml GelRed (Fermentas GmbH, Germany) under UV illumination and photographed. 4.2.4 DNA sequencing of A. marginale msp1a gene PCR products containing only single amplification products of msp1a gene were sequenced at lnqaba Biotech facilities (lnqaba Biotechnical Industries (Pty) Ltd, Pretoria, Gauteng, South Africa). The termination reactions were performed using BigDye VER3.1 (ABI, Life Technologies, CA, USA) according to manufacturer's instructions. The labelled fragments were purified using Zymo research sequencing clean-up kit (Zymo Research, CA, USA) and subsequently analysed on a 3500xl Genetic Analyzer (ABI, Life Technologies, CA, USA). The sequences obtained in this study were submitted to GenBank and provided with accession numbers for msp1a (KC470153-KC470196) gene (Appendix 2). 49 4.2.5 Sequence analysis of A. marginale msp1a gene To identify the msp1a gene sequences obtained in our study the database Nucleotide collection (nr/nt) using Megablast (optimize for highly similar sequences) from the BLAST server was used (Zhang et al., 2000). Protein homology and identity analysis were performed using the multiple-alignment program ClustalW (Thompson et al., 1994). The MSP1 a tandem repeats found in this study were reported previously (Cabezas-Cruz et al., 2013). 4.2.6 Codon based phylogenetic analysis of tandem repeats Codon based alignment was performed using the codon suite server (Schneider et al. , 2005; 2007). Detection of selection pressure on individual codons was calculated using two methods, single likelihood ancestor counting (SLAC) and fixed effects likelihood (FEL) (Pond and Frost. , 2005), used in the Datamonkey webserver (Delport et al., 2010; Pond and Frost. , 2005). Positive and negative selections were assigned to codon where w=dN (non-synonymous substitutions)/dS (synonymous substitutions) ratio was higher or lower than 1 respectively. The reconstruction of the ancestral amino acid sequence was performed using a neighbour joining tree rooted in tandem repeat 83 under the Dayhoff model of substitutions which was estimated to be the best model fitting the actual data. Three reconstruction methods were used: Joint (Pupko et al. , 2000), marginal (Yang et al., 1995) and sample (Nielsen, 2002) which are also used in the Datamonkey webserver. 4.2.7 Amino acid variability, composition and genetic diversity index (GDI) The amino acid variability was calculated in the variability server (Garcia-Boronat et al. , 2008) using the Shannon entropy (H) formula (Shannon, 1948) as fol low: M H = - LP;log2 P; i= I 50 Where Pi is the fraction of residues with a certain type of amino acid i, and M is the number of types of amino acid for a position. The proportion of variable over conserved positions was calculated as the number of positions with more than 0 of Shannon variability divided by the number of positions with 0 Shannon variability. Am ino acids were also classified regarding biochemistry properties, to know: negative charged, positively charged, uncharged-polar and non-polar. For comparison purposes, Shannon variability was additionally calculated in 28 and 43 MSP1 a sequences avai lable in GenBank from Venezuela and USA respectively. To further characterize the genetic diversity of MSP1 a, a genetic diversity index (GDI) was calculated for each A. marginale strain as follow: number of different MSP1 a tandem repeats divided by the total number of tandem repeats per strain . 1 and 0 were considered as maximum and minimum genetic diversity, respectively. Table 5: Observed prevalences of Anap/asma margina/e in different provinces of South Africa South African No. of blood Prevalence of A. No. of msp1a provinces samples collected marginale, sequenced per province msp1a-positive PCR (%) Limpopo 20 13(65) 8 Mpumalanga 21 21 (100) 2 Gauteng 20 20(100) 5 North West 33 24(72) 7 KwaZulu-Natal 55 50(90) 9 Eastern Cape 40 40(100) 3 Western Cape 16 14(87.5) 11 Northern Cape 45 0(0) 0 Free Statea 215 129(60) 29 a Data collected by Mtshali et al. , (2007). 51 4.3 Results and discussion 4.3.1 Molecular evidence of A. marginale prevalence in South Africa Bovine anaplasmosis has been considered to be endemic in South Africa, an assumption based primary on the distribution of the tick vectors (Ndou et al. , 2010) and the sero-prevalence of A. marginale which was determined only in the Free State (Dreyer et al., 1998), Limpopo (Rikhotso et al., 2005) and North West (Ndou et al. , 2010) provinces. Molecular evidence of endemic bovine anaplasmosis has not been reported in most of South Africa. This study was therefore designed to investigate the molecular evidence of A. marginale infection in cattle from Mpumalanga, Gauteng, Eastern Cape, Limpopo, North West, KwaZulu-Natal , Eastern Cape and Western Cape provinces (Table 5). Molecular diagnostic of anaplasmosis was determined previously in Free State province (Mtshali et al., 2007). In the present study, species-specific msp1a primers were used for PCR assays on 250 DNA samples obtained from cattle blood. The results of these PCR studies confirmed that A. marginale is widespread in South Africa, but with a variable prevalence in all the sampled provinces, except for Northern Cape in which no positive samples were detected (Table 5). The absence of A. marginale positive PCR in Northern Cape is not surprising because th is area is considered to be free of tick vectors (Mtshali and Mtshali, 2013) and the prevalence of Babesia spp., another tick-borne pathogen, was also found to be very low in this province (Mtshali and Mtshali , 2013). The provinces with highest prevalence of A. marginale were Mpumalanga, Gauteng, and Eastern Cape with an infection rate of 100%. In the other provinces, A. marginale infections in cattle ranged from 65% to 90% (Table 5 and Figure 7). 52 Differences in the prevalence of A. marginale were not observed between commercial and communal farming systems. The prevalence of A. marginale in cattle from South Africa can be considered high when compared to the prevalence of A. margina/e in cattle from other regions of the world, for example, in Brazil a recent study found 70 % prevalence of A. marginale in a cattle herd (Pohl et al., 2013). Another study, by de la Fuente et al. (2005) showed a range from 25% to 100% of A. marginale prevalence in cattle herds from Italy. Analysis of the prevalence and genetic diversity of A. marginale MSP1 a in different geographic regions constitutes an important step toward development of effective MSP1 a based vaccines because the antigenic composition of the vaccine should contain MSP1 a variants present in different regions of the world . 53 A DDSSSASGQQQESSVSSQSE-ASTSSQLcr- 83 T ............ G.L ... C:;Q •• • ••• S.-- 82 A •....•.•.•...• L •.. DLS .. W .... -- 7 9 T • • • • • • • • P. • • • • Lr::;'f-,/ • DLS • • • . • . . - - 100 T . . . . . . • . . . . . G • L • . . GQ • • . . • • • • - 140 A ............ G.L ... GQ ••..•.• R- 141 T • . • • • • • . • • . . G . L • . . GQ • • • • • • • R- 142 T •..••••.••..•• IaP .• GH.R .... S.- 143 T • . • • G • • . • • . . • • L . . . SQ. RS . • . • . - 144 . . . . . . . . . . . . . . . IaP . . G(JJ) • • • •• s. - 145 G . . . . S • • . • • . . • • L • . . SQ . . . . . . . . - 146 A ..... ~ .......•.... - ........ - 147 T ..... ~ ........... G-.G ...... - 148 T ....• GD •.....•••.. G- •• A .•K .R- 149 T ••.•• @ ••••••••• L.G- ........ - 150 T ..... W .K ........ IG- ..... K .. - 151 AN'. . • . • •.• E . . . . L. . . DQ. . . . .... - 152 T . . . . . . . . . . . . . . L . . . DQ . . . . . . . R - 153 T . . . PE- . . . . . . . . F . . . AQ . • . . . . S . - 154 A . . . . . . . . . . . . . . L . . . DQ . . . . . . S . - 155 AN' • • • • • • • • • • • • • L . . . GQ • • • • • • • • - 158 T .. W. .......... L ...D Q ........ - 160 A • • • • • • • • • • • • • • P • • • • - • A. • • • • AD 161 ~ ............ L •.. DQ •••.•... - Figure 6: Newly reported sequences of MSP1 a tandem repeats. The one-letter amino acid code was used to depict MSP1 a repeat sequences. Dots indicate identical amino acids, and gaps indicate deletion/insertions. The ID of each repeat form was assigned previously in Cabezas-Cruz et al. , (2013). Tandem repeat A was used as a model for amino acid comparison. 54 4.3.2 A. marginale prevalence and msp1a genetic diversity The sequence of A. marginale msp1a was analysed in 44 strains (Table 4). We found 52 different types of tandem repeats among South African strains (including those reported by Mtshali et al. , 2007 from Free State province) from which 23 were described for first time (Figure 6). Using a genetic diversity index (GDI), the genetic diversity per A. marginale strain we described (Table 6) and the average of genetic diversity was calculated per Province. A polynomial correlation (R2=0.7 6) occurred between the GDI and the prevalence of anaplasmosis per province (Figure 7). Interestingly, provinces with 100 % of A. margina/e prevalence were not the provinces with the highest or lowest GDI , rather these provinces were between a range of genetic diversity (0.82 - 0.87). Immune selection and transmission fitness have been suggested to impact genetic diversification of A. marginale (Palmer and Brayton, 2013). Immune pressure in cattle may induce greater genetic diversity in A. marginale populations in order to insure persistent infection. At the same time, the development of antigenic variation was suggested to have a transmission cost (Palmer and Brayton, 2013). A. marginale strains with lower transmission fitness due to high genetic variability would be at risk of elimination from the population (Palmer and Brayton, 2013). Furthermore, high incidence of ticks correlated to increased MSP1 a genetic variability in Argentina (Ruybal et al., 2009). These apparently contradictory aspects may be reconciled if increased tick transmission contributes to greater circulation of A. marginale strains in the cattle population, favouring the interaction between A. margina/e strains and potentially resistant hosts. In this study, the observed range of GDI in which 100% prevalence (an indicator of transmission efficiency) exist may reflect a positive balance between genetic diversity and biological transmission by ticks taking in account that, as shown in Figure 7, all the studied area, except for Northern Cape, in infected by ticks. 55 Notably, most of the new tandem repeats shown in Figure 6 were sequenced from KwaZulu-Natal province which has 90 % prevalence of A. marginale. When KwaZulu- Natal was excluded from the correlation analysis between A. marginale prevalence and GDI (Figure 7), the polynomial regression increase from R2=0.76 to R2=0.98. A possible explanation of this result may be that the tandem repeats diversification found among KwaZulu-Natal MSP1 a provided a degree of transmission fitness to the strains present in that area. Another explanation could be that cattle immunity is increasing in the area, inducing A. marginale diversification for antigenic variation. One interesting question that remains unanswered is how the genetic diversity of A. marginale msp1a emerges in a specific region. 4.3.3 Evolution of msp1a genetic diversity In order to analyse the evolution of msp1a genetic diversity observed in our samples, the tandem repeats (Table 6) were classified as "frequent" (present more than 22 times) or "rare" (present less than 10 times) based on the frequency of their appearance among South African A. marginale strains, including those from Free State province reported by Mtshali et al. (2007). The unique tandem repeats found in this study were classified as rare, being repeated, most of them, one time in only one strain (Figure 6). In contrast, tandem repeats 3, 4, 13, 34, Q and 37 had a high frequency (Table 6) and were also reported in A. marginale strains from Israel (3, 4), South America (4, 13) and Europe (Q). Repeat sequences 34 and 37 were abundant only in South Africa with rare exceptions (as reviewed by Cabezas-Cruz et al., 2013). Considering this, we wanted to test whether the tandem repeats newly described in this study (Figure 6) originated from extant A. marginale MSP1 a tandem repeats or had evolved from a tandem repeat that was lost after tandem repeat differentiation. 56 Table 6: Anaplasma margina/e strains and putative 20 structures of MSP1a tandem repeats A. marginale Province of Structure of msp1a GOia AVE-GDl/STDEV origin strain tandem repeats LP-7 Limpopo 34 159 0.917/0.144 LP-10 Limpopo 27 13 3 36 LP-30 Limpopo 27 13 3 LP-34 Limpopo 34 13 3 38 LP-37 Limpopo 27 13 13 37 0.75 LP-46 Limpopo 3 38 LP-50 Limpopo 34 13 13 0.667 MP-C2 Mpumalanga 34 1 15 37 0.875/0.177 MP-C5 Mpumalanga 15 15 100 83 0.75 NW-C2 North West 27 13 4 4 37 0.8 0.960/0.089 NW-C4 North West 27 13 4 37 NW-C5 North West 82 13 79 4 37 NW-C1-160312 North West 34 13 3 36 38 NW-C4-160312 North West 34 36 38 3 GP-C1 Gauteng 82 13 4 4 37 0.8 0.826/0 .173 GP-C2 Gauteng 34 27 3 38 13 3 38 0.714 GP-C5 Gauteng 3 4 4 4 37 0.6 GP-C11 12105 Gauteng 34 37 GP-C4117105 Gauteng 3 36 38 GP-C7117105 Gauteng 34 13 13 0.667 GP-C1817105 Gauteng 34 13 37 KZN-D KwaZulu-Natal 42 43 25 163 31 0.919/0.128 KZN-F KwaZulu-Natal 4 43 25 31 31 0.8 KZN-K KwaZulu-Natal 27 13 4 4 37 0.8 KZN-Y KwaZulu-Natal 143 144 145 146 KZN-MM KwaZulu-Natal 42 43 25 31 KZN-14 KwaZulu-Natal 142 43 25 31 57 KZN-19 KwaZulu-Natal 141 140 140 0.667 KZN-49 KwaZulu-Natal 147 148 149 150 KZN-51 KwaZulu-Natal 147 EC-22 Eastern Cape 27 13 4 4 37 0.8 0.867/0.115 EC-23 Eastern Cape 151 152 4 4 153 0.8 EC-24 Eastern Cape 27 13 4 WC-4 Western Cape 4 Q Q m 0.75 0.741 /0.286 WC-6 Western Cape 3 4 4 37 0.75 WC-7 Western Cape M M M M 0.25 WC-8 Western Cape 34 4 37 WC-10 Western Cape 154 WC-11 Western Cape 40 Q Q Q Q Q 37 0.429 WC-12 Western Cape 27 13 37 WC-13 Western Cape M Q M Q M 0.4 WC-14 Western Cape 155 36 38 WC-15 Western Cape 161 13 37 4 162 WC-16 Western Cape 34 13 4 13 13 4 37 0.571 More common tandem repeats are highlighted (bold) and mentioned in order of abundance, from Higher to lower: 13,4,37 and 3, 34 have the same frequency, respectively. aGenetic diversity index of MSP1 a (GDI) was calculated as follows: number of different MSP1 a tandem repeats/total number of tandem repeats per strains. 1 is maximum genetic diversity. The average of GDI (AVE-GDI) and standard deviation (STDEV) for each region is shown. 58 120% Eastern Cape ~--~ 100% • • Mpumalanga G) u c • KwaZulu-Natal G) >"' 80% G) North West ~ ~ • • ~ 60% • Free State Limpopo .as, e> ea, 40% 20% y = 7387x4- 2534lx3 + 32490x2 -18449x + 3915.7 R1 =0 .7675 0% '--~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~- 0.7 0.75 0.8 0.85 0.9 0.95 1 GDI Figure 7: Correlation between msp1a genetic diversity and A. marginale prevalence in South Africa. The prevalences of anaplasmosis in different South African provinces were plotted against the average of genetic diversity index (GDI). GDI was calculated for each strain as the number of different tandem repeats divided by the total number of tandem repeats. A polynomial correlation was found between these 2 parameters with R2 = 0.7 6. Provinces with 100% prevalence (Gauteng, Eastern Cape, and Mpumalanga) are in the range of 0.82-0.87 GDI whi le regions having less than 100% prevalence are out of this range. 59 To test this hypothesis we reconstructed the ancestral state (Material and Methods) of the new tandem repeats presented in Figure 6. Surprisingly, the ancestral state of all the new tandem repeats was the tandem repeat 4 (Figure 8), suggesting that all of the newly described tandem repeats from South African A. marginale strains evolved from this tandem repeat. In addition, this evidence suggests that these repeated sequences may constitute a group of recently evolved tandem repeats specific to South Africa and, in fact, these repeats have not been reported elsewhere (Cabezas-Cruz et al., 2013). The mechanism for the generation of genetic diversity of A. marginale could be gene duplication followed by either mutation or mismatch repair as proposed by Palmer and Brayton. (2013). The variab le number of tandem repeats found among MSP1 a isolates suggests that tandem repeat duplication due to homologous recombination may occur in order to give a "substrate" for "testing" competitive advantage of new tandem repeats variants. As demonstrated by this research, new tandem repeats shown in Figure 6 evolved from an initial tandem repeat 4 which served as a template for genetic variation. To test whether tandem repeat 4 evolved to the new forms under selective pressures, the ratio w (see Materials and Methods) was calculated for each codon position of the new tandem repeats from South Africa. We found that the diversification of tandem repeat 4 in South African strains occurred under both positive and negative selective pressures (Table 7 and Figure 8). Surprisingly, the positions evolving under negative selection (Figure 8 negative signs), 8 and 10, were reported before to be present in an immunodominant B-cell epitope present in MSP1 a (Figure 8, first boxed area, Garcia- Garcia et al. , 2004) and the position 25, also evolving under negative selection, was found previously in a neutralization-sensitive epitope (Figure 8, second boxed area; Palmer et al., 1987; Allred et al., 1990). In contrast, one of the positions found to be evolving under positive selection was the amino acid position 20 (Figure 8, arrow) which has been implicated in the binding of MSP1 a to tick cells extract (de la Fuente et al., 2003a). 60 These findings support the hypothesis discussed above that immune escape and tick transmission are both driving forces of A. marginale MSP1 a genetic diversification and suggest that tick transmission and immune escape would be triggers of the observed diversification of tandem repeat 4 in South Africa. This trend of purify ing selection/negative selection (or deletion of the unfit) for sites involved in immune recogn ition by the host was confirmed, in wider frame, by the following data: (i) most of the residues of the MSP1 a immunodominant B-cell epitope (Garcia-Garcia et al., 2004) were deleted in the tandem repeats forms a and 108: the tandem repeat a is widespread in strains from Mexico, Brazil , Venezuela, Argentina and Taiwan and is also present in the most common A. marginale strain of the world which have the tandem repeat composition {a, ~ . ~. ~. r ) (Cabezas-Cruz et al., 2013), and (ii) the first glutamine (Q) of the neutralization-sensitive epitope (Palmer et al., 1987; Allred et al., 1990) is deleted in several tandem repeats {A, D, E, y, L, q> , 5 - 9, 14, 31 , 36, 52, 57, 60 - 66, 69 - 72, 76, 84- 86, 95 - 99, 105, 107, 116-119, 129-131 , 136 - 139) from A. marginale strains reported previously (Cabezas-Cruz et al. , 201 3). As seen in Figure 8, Q was deleted in some of South African new tandem repeats (146, 147, 148, 149 and 150). Collectively, this evidence suggests that purifying selection is most likely one of the mechanisms that A. marginale had evolved to escape immune recognition toward MSP1a. Tandem repeat 4 from MSP1a was also found in A. marginale strains isolated from an anaplasmosis outbreak in Mexico where R. microplus was implicated as tick vector (Almazan et al., 2008), and new tandem repeats were reported also in th is study. It should be interesting to test whether these newly described Mexican MSP1 a variants evolved from tandem repeat 4. We consider these findings to be relevant to the development of MSP1 a based vaccines and suggests that MSP1 a based vaccination should be combined with tick control strategies in order to minimize genetic diversity of A. marginale msp1a. A recent study was reported that combined use of a tick protective antigen with MSP1 a in the same vaccine formulation that was directed toward control both tick infestations and anaplasmosis (Torina et al., 2014). 61 Table 7:Sites that evolved under positive and negative selection in the new tandem repeats from South Africa Codons Methods SLAC Methods FEL Type of selection w(dN/dS p value w(dN/dS p value 1 5.49 0.043 Infinite 0.012 Positive 6 1.58 0.311 Infinite 0.093 Positive 16 1.39 0.371 Infinite 0.196 Positive 20 2.86 0.172 Infinite 0.133 Positive 28 2.22 0.197 Infinite 0.079 Positive 8 -1 .66 0.373 -5.14 0.236 Negative 10 -6.81 0.037 -13.89 0.016 Negative 25 -3.57 0.134 -6.56 0.110 Negative 62 ~ + + -- + + - + -------------1.. --160 ANG1SSASGQQQESSWLsQSDQASTSSQLG 149 TDSSSAGDQQQESSVSSLSG-ASTSSQLG __, .. ·-------145 GDSSSSSGQQQESSVLSQSSQASTSSQLG -.--158 AESSSASGQQQESSVLSQSDQ-STSSQLG __, ...--.... 148 TDSSSAGDQQQESSVSSQSG-ASASSKLR ·-----142 TDSSSASGQQQESSVLPQSGBARTSSQSG .---100 TDSSSASGQQQESGVLSQSGQASTSSQLG ---------....... 161 ADSSSASGQQQESSVLSQSDEASTSSQLG _____. ...,._. ---18421 TDSSSASGQQQESGVLSQSGQASTSSQLR ADSSSASGQQQESSVLSQSDLSSTWSQLG -.--154 ADSSSASGQQQESSVLSQSDQASTSSQSG __, ...--.... 150 TDSSSAGDQKQESSVSSQXG-ASTSSKLG ·-----143 TDSSSGSGQQQESSVLSQSSQARSSSQLG .---151 ANSSSASGQQEESSVLSQSDQASTSSQLG --------~··-r------'--1--41447 TDSSSAGNQQQESSVSSQSG-AGTSSQLG DDSSSASGQQQESSVLPQSGQX>STSSQSG • .---140 ADSSSASGQQQESGVLSQSGQASTSSQLR __, ...--.... 79 TDSSSASGQPQESSVLCVSDLSSTSSQLG I -1 ----153 TDSSPEMGQQQESSVFSQSAQASTSSQSG ·--------155 ANSSSASGQQQESSVLSQSGQASTSSQLG -------------------152 TDSSSASGQQQESSVLSQSDQASTSSQLR 1,_,_ --_-_-_-_-_--_-_-_-_-_--_-_-_-_-_--_-_-_-_1_4 836 ADSSSAGNQQQESSVSSQSE-ASTSSQLG TDSSSASGQQQESGVLSQSGQASTSSQSG ~R 4 : TDSSSASGQQQESSVLSQSG~STSSQLG • Figure 8: Reconstruction of ancestral amino acid sequence and amino acid positions under positive and negative selection. The reconstruction of the ancestral state (TR 4: Tandem repeat 4, sequence marked with *) of the new tandem repeats found in South Africa (Figure 6) was performed using 3 reconstruction methods, namely: joint, marginal, and sample (see 'Materials and methods' for details). Positions that evolve under negative (-) and positive (+) selection are shown (see Table 7). Amino acid at position 20 is indicated (arrow). The residues of the immunodominant 8-cell epitope (Garcia-Garcia et al., 2004) (first box) and the neutralization-sensitive epitope (Palmer et al., 1987; Allred et al., 1990) (second box) are also shown. 63 4.3.4 Amino acid variability and low variable msp1a peptides Genetic variability could also be analysed by calculating the amino acid variability in each position of the tandem repeat. In order to compare the msp1a genetic variability in South Africa with other regions of the world we calculated the amino acid variability for 29 amino acid positions from all the msp1a tandem repeats found in South Africa, USA and Venezuela which were available in GenBank and collected by Cabezas-Cruz et al., 2013. The amino acid variability of msp1a tandem repeats from South Africa can be seen in Figure 9 as compared with that from USA (low) and Venezuela (high). In agreement with th is, the average of amino acid variability was 0.30, 0.42 and 0.7 2 for USA, South Africa and Venezuela, respectively (Figure 9A-C). In comparison the proportion of variable over conserved positions in MSP1a tandem repeats was higher Venezuela (29) compared to USA (0.85) and South Africa (2.2). Epitopes from MSP1 a were found to induce a protective immune response against A. marginale (Santos et al. , 2013). Considering this, using the variability server (Garcia-Boronat et al., 2008), we explored whether some low variable peptides could be found in MSP1 a from South Africa, USA and Venezuela. We observed that the amino acid variability of MSP1a tandem repeats from South Africa and USA supported the existence of low variable peptides (Figure 9) which may be useful in development of peptide-based MSP1 a vaccines, while the high amino variability of MSP1 a tandem repeats from Venezuela do not support for this type of low variable peptide. Interestingly, the low variability peptide in South Africa overlap with the position of the immunodominant B-cell epitope reported previously for MSP1 a (Garcia-Garcia et al., 2004). 64 Venezuela Av.rage of' amino acids ~rlablllty: 0 .72 Proportion ot' vartable/cons..-"9d sit-: 29 3 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 2 4 25 26 27 28 29 •Negative • Positive m Uncharged- polar • Nonpolar a Deleted USA Av.r3ge of' amino acids ~rlablltty: 0.30 3 Proportion of' vartablelcons..-"9d sit-: 0 . 85 ~ 1-i -c: 2 ,.... c: -: ., c: ~ U> 0 ., 2 3 4 5 6 7 8 9 .,0 .,., .,2 .,3 14 ., 5 16 17 18 19 20 21 22 23 2 4 25 26 27 28 29 G Q Q Q E S S V SS Q s* -Negative • Positiv e 1:11 Uncharge d -polar - Nonpolar o Oeleted South Af'rfc.a Av.r3ge of' amino acids ~rlablllty: 0 .42 3 Proportion of' vartable/conser\19d sit-: 2.2 ~ 1i 2 1i > c: -2 ., c: ~ U> 0 ., 2 3 4 5 6 7 8 9 '10 1., 1 2 1 3 14 15 ., 6 17 18 19 2 0 2 1 22 23 2 4 25 26 27 28 29 sis s As o o o o E s slv * •Negativ e • Positive £J Uncha rged- polar • Nonpola r CJ Deleted Figure 9: MSP1a amino acid variability, composition. and low-variable peptides. The figure shows the amino acid variability and composition among the tandem repeats in 3 different countries: Venezuela (A). USA (B), and South Africa (C). Venezuela shows a high proportion of variable/conserved sites and a high average of amino acid variability while South Africa and the USA show middle and lower values, respectively. Different colours in columns depict different biochemical properties in the amino acid composition: negative (green). positive (red), uncharged- polar (beige), and non-polar (blue); proportion of deleted positions is shown in yellow. Consensus sequences of low-variable peptides are shown (*) for the USA and South Africa. The region of the immunodominant B-cell epitope from A. marginale (Garcia-Garcia et al., 2004) is boxed in the low- variable peptide from South Africa. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.) 65 CHAPTER 5 STRUCTURAL AND PHYLOGENETIC ANALYSES OF ANAPLASMA MARG/NALE INFECTING CATTLE IN SOUTH AFRICA USING msp1a AND msp4 GENES 5.1 Introduction Phylogenetic analysis establishes the relationships between genes or gene fragments, by inferring the common history of the genes or their fragments. To achieve this, it is essential that homologous sites be compared with each other (positional homology) (Nei and Kumar, 2000). Phylogenetic studies have been previously conducted to resolve and understand the Anaplasma species and how these species relate to one another (de la Fuente et al. , 2007a). The phylogenetic relationship and evolution of A. marginale isolates is important for understanding the biology and the possibilities for control of anaplasmosis (Kocan et al., 2002). Recent research has focused on MSPs that are involved in interactions with vertebrate and invertebrate host cells (McGarey and Allred, 1994; McGarey et al. , 1994 de la Fuente et al., 2001 b) and have been used to elucidate the phylogeographic patterns of A. marginale (Kano et al., 2002; Lew et al., 2002; de la Fuente et al., 2007a). MSP1 a and MSP4 genes were used previously, for phylogenetic analysis that was found to be involved in interactions with both vertebrate and invertebrate hosts (McGarey and Allred, 1994; McGarey et al. , 1994; de la Fuente et al. , 2001b). Therefore, these MSPs are also likely to evolve more rapidly than other nuclear genes because they are subjected to selective pressures exerted by host immune systems. 66 Both msp1a and msp4 genes have been found to be stable genetic markers during multipl ication of A. marginale (de la Fuente et al., 2001a; Bowie et al., 2002). MSP1a was also shown to effect transmission of A. marginale by ticks (de la Fuente et al. , 2001 a; Kocan et al. , 2010). These genes were reported to be more stable for phylogenetic studies because the MSPs do not appear to undergo antigenic variation in cattle or ticks (Bowie et al., 2002). MSP1 a, encoded by msp1a, has been reported to be an adhesin for bovine erythrocytes and tick cells and to effect adhesin, infection, and transmission of A. marginale by ticks of the genus Dermacentor (McGarey and Allred , 1994; McGarey et al., 1994; de la Fuente et al., 2001 c). Although the specific function of MSP4 is not known, previous analysis of the msp4 gene from A. marginale isolates demonstrated that its sequence varies sufficiently among isolates to support its use in phylogeographic studies (de la Fuente et al. , 2003a). In South Africa, phylogenetic profile of A. marginale infecting cattle has not been determined meaning that the genetic relationship between South Africa A. marginale isolates and those in other African countries is unknown. Therefore, the aim of this study is to establish the phylogenetic profile of A. marginale infecting South Africa cattle using msp1a and msp4 as marker genes. 5.2 Materials and methods 5.2.1 DNA extraction The genomic DNA was extracted from cattle blood samples using ZR Genomic DNA TM Tissue Miniprep (Zymo Research, CA, USA). Total volume of blood sample was adjusted in a microcentrifuge tube with water to 100 µI prior to adding 95 µI of 2x Digestion buffer and 5 µI of Proteinase K. The mixture was thoroughly mixed and then incubated at 55 °C for 20 minutes. A volume of 700 µI of genomic lysis buffer was added to the tube and mixed thoroughly by vortexing. 67 The mixture was carefully transferred to a Zymo-Spin ™ llC column in a collection tube and centrifuged at 10,000 x g for one minute. A volume of 200 µI of DNA pre-wash buffer was added to the spin column in a new collection tube and centrifuged at 10,000 x g for one minute. A volume of 400 µI of g-DNA wash buffer was added to the spin column and centrifuged at 10, 000 x g for one minute. The spin columns were carefully transferred to a clean 1.5 ml microcentrifuge tube and 100 µI of DNA Elution buffer added to the spin column. The microcentrifuge tube was incubated for 5 minutes at room temperature, and then centrifuged at 8000 rpm for 30 seconds to elute the DNA. The eluted DNA was stored at -20 °C until used for studies. The concentration of DNA was measured using a NanoDrop® ND-1000 (NanoDrop Technologies Inc., Wilmington, USA). 5.2.2 Set of primers used for amplification of msp1a and msp4 genes Table 8: Oligonucleotide sequence of primers used in this study and their references Primer Oligonucleotide (5'-3') Reference designation 1733F TGTGCTTATGGCAGACATTTCC Lew et al., 2002 2957R AAACCTTGTAGCCCCAACTTATCC Lew et al., 2002 MSP45F GGGAGCTCCTATGAATTACAGAGAATTGTTTAC de la Fuente et al., 2003a MSP45R AAGCTCGAGGCTGAACAGGAATCTTGCTCCAAG de la Fuente et al. , 2003a 68 5.2.3 Anaplasma marginale species-specific PCR of the msp1a gene Specific set of primers were used to amplify msp1a gene (Table 8). Primers were synthesized and supplied by lnqaba Biotech (lnqaba Biotechnical Industries (Pty) Ltd, Pretoria, Gauteng, South Africa). The PCR reagents mixture consisted of total volume of 25 µI of Thermo Scientific DreamTaq Green PCR Master Mix (2X), 10 µM of forward and reverse primers, 1 µg Template DNA in 0.2 ml PCR tubes. The PCR mixture was then amplified by DNA thermocycler (Bio-Rad T100™ DNA Thermal Cycler, Bio-Rad Laboratories, Johannesburg, South Africa). The amplification process involved an initial denaturation at 94 °C for 3 minutes followed by 40 cycles of denaturation at 94 °C for 1 minute at 94 °C, annealing at 65 °C for 1 minute, and an extension at 72 °C for 2 minutes. A cycle of a final extension was done at 72 °C for 7 minutes. The amplified products were then separated by electrophoresis using 1% agarose gel immersed in TBE buffer (89 mM Tris-Borate, 2 mM EDTA, pH 8). One kilo-base pair ladder marker was used to determine the size of the PCR products after staining with 0.5 µg/ml GelRed (Life Sciences, Fermentas GmbH, Germany) and visualization under UV illumination. Gel documentation was done and taking a photograph using a camera. 5.2.4 Anaplasma. marginale species-specific PCR of the msp4 gene The msp4 gene was amplified from 5 µI per 20 µI reaction volume using 10 pmol of each primer (Table 7) (2x Phusion Flash PCR Master Mix) (Thermo Fisher Scientific, USA). Reactions were performed in a BIO-RAD T100™ DNA thermal cycler (Bio-Rad Laboratories, Johannesburg, South Africa). The amplification cycles following an initial denaturation at 98 °C for 3 minutes for 1 cycle then consisted of 3-step protocol of a denaturation step of 98 °C for 10 seconds, annealing at 63 °C for 30 second and extension at 72 °C for 30 seconds for 5 cycles, 98 °C for 1O seconds, annealing at 65 °C for 30 seconds and extension at 72 °C for 30 seconds for 5 cycles, 98 °C for 10 seconds, annealing at 67 °C for 30 seconds and extension at 72 °C for 30 seconds for 20 cycles, with a final cycle at 72 °C for 7 minutes. 69 The ampl ified products were then separated by electrophoresis using 1 %agarose gel immersed in TBE (89 mm Tris-Borate, 2 mm EDTA, pH 8). One kilo-base pair ladder marker was used to determine the size of the PCR products after staining with 0.5 µg/ml GelRed (Life Sciences, Fermentas GmbH, Germany) and visualization under UV illumination. Gel documentation was done and taking a photograph using a camera (Figure 10). 70 M 2 3 4 5 6 7 8 1000 900 800 700 600 500 400 300 200 100 Figure 10: Gel image of 1% agarose gel electrophoresis of PCR amplification products obtained from A. marginale from Western Cape Province in South Africa using primer MSP45F and MSP45R. Lane M: GeneRuler™ 1 Kb DNA ladder, ready-to-use; Lane 1- 7 indicate positive results and lane 8 indicate negative control. 71 5.2.5 DNA sequence alignment and phylogenetic analysis All A. marginale msp1a and msp4 gene sequences and GenBank accessions utilised here are listed in Appendix 2, 3 and 4. The obtained sequences were edited using Chromas V1 .49 software and BioEdit (Tom Hall Ibis Biosciences, Carlsbad , USA). Nucleotides were coded as unordered, discrete characters with five possible character states: A, C, G, T or N and gaps were coded as missing data. The percent identity between the tandem repeat in position 1 (R1) and the other tandem repeats was determined using CLUSTALW. The tandem repeats secondary structure was predicted using the PSIPRED server (Buchan et al., 2010). The sequences obtained in this study were compared with available sequences from the NCBI database using the Basic Local Alignment Search Tool (BLAST) (http://www.ncbi.nlm.nih.gov/BLAST/). The sequences were submitted to GenBank (Appendix 3). Multiple sequence alignment was performed using ClustalW 1.6 program (Thompson et al., 1994). Phylogenetic analysis was conducted in MEGA 5 (Tamura et al. , 2007) using Neighbour joining (NJ) and Maximum Likelihood (ML) algorithm (Saitou and Nei, 1987). The evolutionary distances were computed using with Kimura 2- parameters correction (Kimura, 1980) with bootstrap analysis of 1000 replicates. Sequences of A. marginale msp4 reported previously from West, East and North Africa, North and South America, Europe, Asia and Australia were obtained from GenBank and used for the msp4 gene phylogenetic analysis of South African isolates sequences (Appendix 3 and 4). The A. ovis and A. phagocytophilum (GenBank accession number AF393742 and AF530197 respectively) were used as outgroups. 72 A total of 5 phylogenetic trees were constructed to determine the relatedness of South African A. marginale strains with strains from West, East and North Africa, North and South America, Europe, Asia and Australia (Appendix 4). One phylogenetic tree was constructed based on the msp1a gene sequences using the maximum likelihood method (Figure 12). Two phylogenetic trees were constructed based on msp4 gene sequences using neighbor- joining method and the other two using maximum-likelihood method (Figure 13a and 14a). One phylogenetic tree for each neighbor-joining and maximum likel ihood method was constructed from South African province isolates and the other with South African province isolates together with the sequences obtained from West, East and North Africa, North and South America, Europe, Asia and Australia the obtained from the GenBank (Figure 13a and 13b) (Figure 14a and 14b) (Appendix 4). Anaplasma ovis and A. phagocytophilum were used as outgroups for the construction of phylogenetic tree based on msp4 gene sequences. 5.3 Results Previous studies revealed that the gene and protein sequences of msp1a and msp4 nucleotide sequences have been used to infer the phylogenetic relationships among Oklahoma and New World isolates from Argentina, Brazil, Mexico and the United States. All the 11 A. marginale isolates collected from Oklahoma had diffe rent msp 1a sequences but identical to msp4 sequences (de la Fuente et al., 2003a). Phylogenetic studies of South African isolates using msp4 gene sequences demonstrated that A. marginale isolates had a 100% and 99% identity with the sequences from West, East and North Africa, North and South America, Europe, Asia and Australia sequences available in the GenBank (Appendix 4). 73 5.3.1 Phylogenetic analysis of A. marginale isolates using msp1a gene 5.3.1.1 MSP1 a sequence analysis The msp1a sequences of 44 A. marginale isolates were analysed as shown in Table 6. Differences were found in the tandem repeat sequences and structure of msp1a genes among the isolates included in th is study. The msp1a tandem repeats gave 23 new sequences with amino acid changes (Figure 6) resulting in identification of forty four different strains of A. marginale (Table 6). The strain containing the tandem repeat structure 27/13/4/4/37 was found twice as KZN-K (Table 6). Tandem repeats 3, 13, 27, 34 and 37 were common in the study areas in all provinces and tandem repeat 42 and 43 were highly represented in KwaZulu-Natal (Table 6). A pattern of proportional decrease in the % of homology of R1 with the subsequent tandem repeats (R1 + n) was obtained. The % of homology between R1 and the tandem repeat in position (R1 + n) will have a decrease proportional to the % of homology between R1 and the tandem repeat in position (R1 + 1) (Figure 11 ). Differential pattern of proportional decrease in amino acid homology was found in KwaZulu-Natal (Figure 11 ). On the other hand, the percent of homology differed among the tandem repeats in the provinces in Figure 11 a (approximately 90% and 70% respectively). 74 South African Identity among tandem repeats A provinces with this pattern 96% R1 Limpopo Mpumalanga 93% North West Eastern Cape 11 89% Gauteng Western Cape B 89% R1 KwaZulu-Natal 7<:1/'o 11 42% Figure 11: Proportional identity between tandem repeat in position R 1 and R 1+ n. This pattern was used to suggest a flow of transmission/migration from Eastern Cape to KwaZulu-Natal. The strain containing the tandem repeat structure 27/1 3/4/4/37 was found twice, once in KwaZulu-Natal and then again in the Eastern Cape (Figure 12). Taking in account that this strain does not follow the pattern of proportional decrease found in KwaZulu-Natal, it was concluded that th is strain was introduced from Eastern Cape and this conclusion was in agreement with the maximum likelihood phylogenetic analysis of strains found in KwaZulu-Natal, where KZN-K (27/13/4/4/37) was the only one apart the cluster formed by the other strains from that province (Figure 12). 75 Table 9: Anaplasma marginale strains and putative 20 structures of MSP1a tandem repeats A. marginale Province of Structure of msp1a MSP1a tandem origin strain tandem repeats repeats 20 structure LP-7 Limpopo 34 15 (a-a) (a-a) LP-10 Limpopo 27 13 3 36 (a-a)(a-a)(a-c)(c-a) LP-30 Limpopo 27 13 3 (a-a)(a-a)(a-c) LP-34 Limpopo 34 13 3 38 (a-a)(a-a)(a-c)(a-c) LP-37 Limpopo 27 13 13 37 (a-a)(a-a)(a-a)(a-c) LP-46 Limpopo 3 38 (a-c)(a-c) LP-SO Limpopo 34 13 13 (a-a)(a-a)(a-a) MP-C2 Mpumalanga 34 13 1S8 37 (a-a)( a-a) (a-a)(a-c) MP-CS Mpumalanga 1S 1S 100 83 (a-c)(a-c)(a-c)( a-c) NW-C2 North West 27 13 4 4 37 (a-a)( a-a)( a-13)( a-13)( a-c) NW-C4 North West 27 13 4 37 (a-a)( a-a)( a-13 )( a-c ) NW-CS North West 82 13 79 4 37 ( a-c)( a-a)( a-a)(a-13 )( a-c) NW-C 1-16031 2 North West 3 13 3 36 38 (a-a)( a-a)( a-c)( c-a )(a-c) NW-C4-160312 North West 34 36 38 3 (a-a )( c-a)(a-c)( a-c) GP-C1 Gauteng 82 13 4 4 37 ( a-c )(a-a)( a-13 )( a-13 )( a-c) GP-C2 Gauteng 34 27 3 38 13 3 38 (a-a )(a-a)( a-c)(a-c)( a-a) (a -c)( a-c) GP-CS Gauteng 3 4 4 4 37 (a -c )( a-13 )( a-13 )( a-13 )(a -c ) GP-C111210S Gauteng 34 37 (a-a)(a-c) GP-C411710S Gauteng 3 36 38 (a-c)(c-a)(a-c) GP-C711710S Gauteng 34 13 13 (a-a)(a-c)(a-c ) GP-C18171 OS Gauteng 34 13 37 (a-a)(a-c)(a-c) KZN-D KwaZulu-Natal 42 43 2 1 3 31 (a-c)(a-c)(a-c) (a-a)(a-a) KZN-F KwaZulu-Natal 4 43 2S 31 31 (a-c)( a-c)( a-c)( a-a)( a-a) KZN-K KwaZulu-Natal 27 13 4 37 (a-a) (a-a) (a-13)(a-13)(a-c) KZN-Y KwaZulu-Natal 143 144 14S 146 ( aa)( a-c )( a-c)( a-a ) KZN-MM KwaZulu-Natal 42 43 2S 31 (a-c)(a-c)(a -c)( a-a) KZN-14 KwaZulu-Natal 142 43 2S 31 (c)(a-c)(a-c)(a-a) KZN-19 KwaZulu-Natal 141 140 140 (a-c)(l3-c)(l3-c) KZN-49 KwaZulu-Natal 147 148 149 1 0 (a-c)(a-c)(a-c)(a-c) KZN-S1 KwaZulu-Natal 147 (a-c) EC-22 Eastern Cape 27 13 4 3 (a-a )( a-a)( a-13)( a -13)( a-c) EC-23 Eastern Cape 1 S1 1 2 4 4 1S3 (o -o )(o-a-a )(o -13)( o -13 )( oo) EC-24 Eastern Cape 27 13 4 (a-a)(a-a) (a-13) WC-4 Western Cape 4 Q Q m (o-a )(a-a)(o-a)(a-c) WC-Q Western Cape 3 4 4 37 (a-c)( a-13)(a-13)(a-c) WC-7 Western Cape M M M M (a-c)( a-c)(a -c)(a -c ) 76 WC-8 Western Cape 34 4 37 (a-a)(a-~)(a-c) WC-10 Western Cape 154 (a-c) WC-11 Western Cape 40 Q Q Q Q Q 37 (a-a)(a-a)(a-a)(a-a )(a-a) (a-a)(a-c) WC-12 Western Cape 27 13 37 (a-a)(a-a)(a-c) WC-13 Western Cape M Q M Q M (a-c)(a-a)(a-c)(a-o) (a-c) WC-14 Western Cape 155 36 38 (aa)(c-a)(a-c) WC-15 Western Cape 161 13 37 4 162 (a-a)( a-a)( a-c)( a-~ )(a-a) WC-16 Western Cape 34 13 4 13 13 4 37 (a-o)(a-a )(a-~)( a-a)( a-a) (a-~)( a-c) The A. marginale strains detected in this study are presented as msp1a tandem repeats organization. Limpopo (LP), Mpumalanga (MP), North West (NW), Gauteng (GP), KwaZulu-Natal (KZN), Eastern Cape (EC) and Western Cape (WC). The code-numbers for tandem repeats are shown in Figure 2. The putative secondary structure of every tandem repeat is shown as two separated alpha helices (a-a), one long alpha helix (aa), beta strand (13) and coiled structure (c). The pattern found in LP, MP, NW, GP, KZN, EC and WC provinces is listed (Figure 11 ); reductions of approximately 3% were observed among A. marginale msp1a tandem repeats. The A. marginale strain NW-C4 served as a model (A). Sequences from KwaZulu-Natal did not follow this pattern and reductions of approximately 10% and 40% were observed. The A. marginale strain KZN-14 served as a model (B). 77 A . marutnale KZM-51 0 .06 Figure 12: Maximum- likelihood phylogenetic tree based on the MSP1 a from the strains present in KwaZulu-Natal and North West provinces. In this tree the strain A. marginale KZN-K (red dot) was in the cluster from North West sequences (blue circle). Bootstrap values are shown as a percentage in the internal branch. Only bootstrap values higher than 50% were shown. NJ phylogenetic analysis gave similar results (Data not shown). 78 5.3.2 Phylogenetic analysis using msp4 gene sequences The msp4 gene of A. marginale isolates from Limpopo, Mpumalanga, North West, Gauteng, KwaZulu-Natal , Eastern Cape, Northern Cape and Western Cape provinces of South Africa were amplified by PCR and then sequenced (Figure 10). Positive results were not obtained from Northern Cape samples. Two samples from each province were randomly selected for phylogenetic analysis. Four phylogenetic trees were each constructed for A. marginale isolates randomly selected from the LP, MP, GP, NW, GP, KZN, EC and WC provinces of South Africa (Figure 13a and 14a) (Appendix 3), and those from the other countries in Africa and around the world (Figure 13b and 14b) (Appendix 4). Anaplasma ovis and A. phagocytophilum were included as outgroup taxa for the phylogenetic analysis based on the msp4 gene. 5.3.3 Phylogenetic analysis of msp4 gene using neighbor -joining phylogenetic tree A complete msp4 gene was obtained after sequencing using the primers designed previously to amplify msp4 gene (Appendix 3). The A. marginale msp4 gene was amplified and sequenced isolates infecting cattle from LP, MP, NW, GP, KZN, EC and WC and the results revealed that this nucleotide differed in length and identity. Anaplasma marginale isolates from LP, GP, NW, KZN and WC had seven different nucleotides from MP and EC isolates at position 163, 197, 223, 239, 251 , 281 and 725 with the following changes A to G, G to A, C to A, G to A, A to G, G to A and G to A respectively (Table 10). These changes gave to two clades from South African isolates and were classified as South African first clade and South African second clade. The first South African clade consists of isolates from LP, GP, NW, KZN, and WC and the second South African clade consist of isolates from MP and EC (Figure 13a, 13b, 14a and 14b). The first clade had 99% bootstrap support and was the same as the second clade, thus demonstrating that South African A. marginale isolates are more closely genetically related. 79 Table 10: Differences in the A. marginale msp4 nucleotide sequences of LP,NW,GP,KZN,WC and MP, EC isolates of South Africa Nucleotide position Isolates 163 197 223 239 251 281 725 LP A G c G A G G NW A G c G A G G GP A G c G A G G KZN A G c G A G G WC A G c G A G G MP G A A A G A A EC G A A A G A A In most of the A. marginale msp4 isolates that have been sequenced, a 849 bp nucleotides with coding sequences were amplified. 80 50 Figure 13a: Neighbor-joining phylogenetic tree of A. marginale msp4 gene sequences from strains identified in cattle in South Africa. The phylogenetic tree was implemented in the MEGA5 (Tamura et al. , 2011 ). Bootstrap analysis was conducted with 1000 replicates. The Gen Bank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. 81 AF428084 1 Mexico AF428083 1 Mexico 57 AF428085 1 Mexico Southern hemisphere AY283190 1 Brazil 4 AY191827 1 Puerto Rico AF428086. 1 Argentina 51 '------AY787172 1 lsraeh round 65 o-- ----- -AY456002 1 SP7-Spa1n '----- - --- - JN572Q28 1 GZB.China '-- --------AY786994.1 Israeli tailed AY127076 1 S!Jllwater 2 AY010253.1 Okeechobee 56 AY127073.1 O klahoma City '----+--- AY010254 1 V1rg1n1a 100 68 AY0 10251.1 M1&&1ss1ppt EU315782.1 50 (G16}-Hungary Northern hemisphere S3 EU315783 1 51 (G18}-Hungary AY702921 .1 na1y 47 .-----s-2 t---- AY702917 1 na1y 6 AY786993. 1 Israeli non-tailed 100 AF428081 .1 USA 55 AY127065.1 Oregon AY010248.1 Calllomia S• AY127077 1 New Castle AY127075.1 S tillwater 1 A Y 127068. 1 Glencoe 2 '--- -------- - ----· AY706389 Anaplasma phagocy1ophilum '-------------------• AF39374 2 .1 Anaplas ma ollis .--------- ------ - AY;>83191 1 Brazil 5 A Y666007. 1 2 .4-Zimbabwe ~-7-1 o----AY666006.1 1 6-Zimbabwe AY666011 .1 5 .9-Zimbabwe 100 '------AY666004.1 661Kan-Kenya 90 AY666003.1 G38-Australoa SS A Y666002.1 F72-Austraha .---- -· KF758867 MP-C15 t----· KF758869 MP-C18 81 • KF758845 EC-25A Southern hemisphere • KF758847 EC-33A • KF758848 G P -C33 • KF758852 GP-C56 100 • KF758856 KZN-C29 • KF758857 KZN-C31 • KF758880 NW-B7 • KF758882 NW-88 • KF758904 W C-C18 • KF758905 WC-C20 • KF758931 LP46 • KF758932 LP49 Figure 13b: Neighbor-joining phylogenetic tree of A. marginale msp4 gene sequences from strains identified in cattle in South Africa represented with a green dot and blue dot representing an outgroup (A. ovis and A phagocytophilum) including isolates from East and North Africa, North and South America, Europe and Australia. The phylogenetic tree was implemented in the MEGA5 (Tamura et al., 2011 ). Bootstrap analysis was conducted with 1000 replicates. The GenBank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. Only bootstrap values higher than 50% are shown. 82 5.3.4 The phylogenetic analysis of msp4 gene using maximum-likelihood phylogenetic tree Observations made from the maximum-likelihood phylogenetic trees in Figure 13a further verify the results of the neighbor-joining phylogenetic tree. The only difference was that in the second clade the number of bootstrap support significantly decreased. In Figure 13b, the difference was noted in the first and second clade when isolates from other countries were included with a bootstrap support of 97, respectively. However, the South African strains fell into a separate cluster from the rest of the world and other African countries. 83 Figure 14a: Maximum-likelihood phylogenetic tree of A. marginale msp4 gene sequences from strains identified in cattle in South Africa. The phylogenetic tree was implemented in the MEGA5 (Tamura et al. , 2011 ). Bootstrap analysis was conducted with 1000 replicates. The Gen Bank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. 84 AF428085 1 Mexico } AF428084 1 Mexico 54 Av2s3100. 1 B razil 4 Southern hemisphere AY 191827. 1 Puerto Rico 6 1 AF428086. 1 Argentina AF428083. 1 Mexico SB A Y 787172. 1 Israeli round JN572928. 1 GZB-China AY456002. 1 SP7-Spaln AY786994. 1 Israeli tailed EU315782. 1 50 (G16)-Hungary BB EU315783. 1 51 (G18)-Hungary 6 • A Y 127076. 1 S tillwater 2 A Y 010253. 1 Okeechobee 9 3 A Y 127073. 1 Oklahoma City AY010254. 1 Virginia 63 Northern hemisphere A Y010251 . 1 M ississippi AY786993 1 lsn1eli non-tailed AY702921 . 1 Italy 47 AY702917. 1 Italy 6 AF428081 . 1 USA 77 AY127065. 1 Oregon AY010248. 1 California 93 AY127077. 1 New Castle AY127075. 1 Stillwater 1 A Y 127068. 1 Glencoe 2 e AF393742. 1 Anaplasma o...;s e AY706389 Anaplasma phagocytophllum AY283191 . 1 Brazil 5 A Y 666007. 1 2 .4-Zlmbabwe 68 AY666006. 1 1 .6-Zlmbabwe 9 3 A Y666011 . 1 5 .9-Zlmbabwe 92 AY666004. 1 661Kari-Kenya A Y 666003. 1 G38-Australia 9B AY666002. 1 F72-Australia • KF758867 MP-C15 98 • KF758869 MP-C18 75 • KF758845 EC-25A • KF758847 EC-33A Southern hemisphere • KF758848 GP-C33 e KF758852 GP-C56 9B • KF758856 KZN-C29 • KF758857 KZN-C31 • KF758880 NW-87 98 • KF758882 NW-88 • KF758904 WC-C 18 • KF758905 WC-C20 • KF758931 LP46 • KF758932 LP49 Figure 14b: Maximum-likelihood phylogenetic tree of A.marginale msp4 gene sequences from strains identified in cattle in South Africa represented with a green dot and blue dot representing an outgroup including isolates from East and North Africa, North and South America, Europe, Asia and Australia. The phylogenetic tree was implemented in the MEGA5 (Tamura et al., 2011 ). Bootstrap analysis was conducted with 1000 replicates. The GenBank accession numbers of the respective sequences used for the phylogenetic analysis are indicated at the beginning of the name of each sequence. Only bootstrap values hiqher than 50% are shown. 85 5.4 Discussion The use of molecular data for inferring phylogenetic trees has now gained considerable interest among biologists of different disciplines, and is often used in addition to morphological data to study relationships in further detail. The number of repeats found in msp 1a has been used to characterize geographic isolates of A. marginale (Allred et al. , 1990; de la Fuente et al., 2002a). In a previous study of A. marginale isolates from the United States, de la Fuente and colleagues (2003a) concluded that msp1a is not a good marker for the characterization of A. marginale geographic isolates and suggested that the genetic heterogeneity observed among A. marginale isolates of Oklahoma could be better characterized by use of msp4 gene and protein sequences. Furthermore, concurrent analysis of msp1a and msp4 gene sequences provided phylogeographic information (de la Fuente et al., 2003a). On a broader geographic scale, analysis of New World isolates demonstrated that msp4 sequences, but not msp1a DNA or protein sequences, provide phylogeographic information (de la Fuente et al., 2007a). These results, together with the finding that multiple msp1a genotypes circulate in some geographic regions (e.g., an area of endemicity in Oregon; (Palmer et al. , 2001 ), questioned the use of msp1a sequences for identification of geographic isolates of A. marginale and suggested that msp 1a sequences may evolve more rapidly (de la Fuente et al., 2007a). Overall , these results suggested that msp1 a genotypes most likely reflect the history of cattle movement rather than the geographic distribution of A. marginale isolates. Recent results have shown that multiple A. marginale genotypes were maintained within a herd in an area of endemicity by independent transmission events and that infection with more than one genotype per host may have been prevented by a mechanism of infection exclusion (Palmer et al., 2001 ; de la Fuente et al. , 2007a; Pohl et al., 2013; Cabezas-Cruz et al. , 2013; Mutshembele et al., 2014). 86 Therefore, if cattle persistently infected with a new A. marginale genotype were introduced to a herd, these genotypes could become establ ished by mechanical and/or biological transmission to susceptible cattle. In regions with few introductions of A. marginale isolates such as Australia, genotypic variation was found to be minimal (Bock and de Vos, 2001 ; Lew et al., 2002). In regions like Oklahoma, where the movement of cattle has been extensive, a highly heterogeneous population of A. marginale isolates would be expected (de la Fuente et al., 2007a). In summary, as demonstrated in previous studies (de la Fuente et al., 2005c), the msp4 sequences of A. marginale provided sufficient variation to provide phylogeographic relationships of A. marginale isolates on a geographic scales. In contrast to the resu lts obtained with msp4, DNA and deduced amino acid sequences of msp1a failed to resolve phylogeographic patterns of A. marginale isolates and suggested that cattle movement and the maintenance of different genotypes by independent transmission events explain the genetic heterogeneity observed among isolates of A. marginale in Oklahoma and, possibly, in other regions of endemicity (de la Fuente et al. , 2007a). 5.4.1 Phylogenetic analysis of A. marginale using msp1a gene The tandem repeat region of msp 1a is useful as a genetic marker for analysis of A. marginale evolutionary and diversity in endemic areas (de la Fuente et al., 2007a). As shown previously, A. marginale is endemic in South Africa (de Waal, 2000). This study used a sensitive and specific msp 1a PCR primer pair (Lew et al., 2002) to study the epidemiology of A. marginale in 8 different provinces of South Africa as well as the structural analysis of msp1a. As was found in previous studies (Pohl et al., 2013; Cabezas-Cruz et al. , 2013; Mutshembele et al., 2014) the msp1a sequences of A. marginale isolates from South Africa contained a variable number of tandem repeats in the amino-terminal region of the protein, while the remainder of the protein was highly conserved. 87 Newly described tandem repeats from South African isolates were reported for first time in this study (Figure 6) suggesting strains that are unique to this geographical region. The presence of different msp1 a genotypes identified in different regions (Cabezas- Cruz et al., 2013) suggest that msp 1a sequences, although conserved during multiplication of the parasite in the bovine host and after transmission by ticks (Palmer et al., 2001 ; Bowie et al. , 2002) are very variable and may contribute to the definition of genotype variations within cattle. This heterogeneity among msp 1a gene sequences could be explained by the tandem repeats within a region of high mutability, which would be supported by the frequency of variable amino acid positions within geographical strain that are higher in this region than in the rest of the protein (Bowie et al., 2002). Nonetheless analysis of this diversity should take into account cattle movement and the consecutive multiple introductions of A. marginale strains in a given region (de la Fuente et al., 2010) which could be a non-evolutive mechanism that contributes to the local genetic variability. Despite the fact that no geographic phylogenetic relationship was found using A. marginale msp1a, the percent of identity among msp1a tandem repeats in two different regions differed in this study, which suggest that different genetic mechanisms may have been involved in the msp1a amino acid variability. Using this phylogenetic pattern South African A. marginale strains EC22 and NW-C2 (27/13/4/4/37) were likely introduced in KwaZulu-Natal from Eastern Cape or North West provinces. This supports the evidence that the strains do not fall in the phylogenetic cluster formed by the other sequences from KwaZulu-Natal and notably, these strains contained tandem repeat number 4 with primarily 11-strand as secondary structure. Recently, the presence of 11- strand in msp1a tandem repeats was found to be phylogenetically correlated with non-tick transmission phenotype of A. marginale (Cabezas-Cruz et al., 2013). 88 This is in contrast with the predominance of a-helices in South African strains that correlate with tick transmission phenotype of A. marginale (Cabezas-Cruz et al., 2013). Several tick species have been incriminated in A. marginale transmission in South Africa (Marufu , 2014) with mechanical and transplacental transmission being reported (Aubry and Geale, 2011 ), which may likely also contribute to transmission of South African EC22-(27/13/4/4/37) strain. Anaplasmosis is widespread in KwaZulu-Natal province (du Plessis et al., 1994) and some of the tandem repeats found in our study has also been reported in other regions of the world (Videtto et al., 2006; de la Fuente et al. , 2007a; Mtshali et al., 2007). Although tandem repeats including 25, 31 , 42 and 43 are common in A. marginale isolates in KwaZulu-Natal , a higher proportion of new South African tandem repeats were found in this study. This data suggested that conditions in this region promoted high genetic variability of A. marginale, a finding in agreement with the discovery that tandem repeats in this province had the lowest amino acid identity. High infection rates by A. marginale-isolates were observed in Mpumalanga, Gauteng and Eastern Cape. The high prevalence is attributed to heavy-tick infestation previously reported in the areas by Horak et al (1991) , Walker (1991) and Coetzer et al (1994). In particular R. decoloratus, R. microplus and R. evertsi evertsi have been reported in these regions and some of these species transmit A. marginale. The 20 structure of the msp1a tandem repeats in these provinces was predominantly a- helix, which would be expected in areas heavily infested by these. In contrast, the lack of A. marginale in Northern Cape would be expected in areas where conditions and intensive tick control programs contributed to low tick populations. Amino acid variability in individual positions along the 29 amino acids of all strains suggested that this position is the most variable among isolates worldwide (Cabezas-Cruz et al. , 2013). In addition, th is position has a high level of negatively-charged amino acids, a region which is associated with transmissibility by ticks (Mutshembele et al. , 2014). 89 This study has used structural epidemiology to establ ish for the fi rst time A. marginale msp1a structural patterns for different geographic regions. Putative secondary structure, identity, and phylogenetic analysis of the msp1a tandem repeats were used to infer movement of A. marginale strains and regional specific properties of A. marginale msp1a. Therefore, the find ings of the present study will form a basis for futu re research where the epidemiology of A. marginale can be correlated with the structure of msp1a and tick transmissibility. 5.4.2 Phylogenetic analysis of A. marginale using msp4 gene The analysis of msp4 gene, which has provided evolutionary information about geographically distinct A. marginale strains (de la Fuente et al., 2007a), was used in this study for phylogenetic analysis for isolates infecting cattle from Limpopo, Mpumalanga, North West, Gauteng, KwaZulu-Natal, Eastern Cape and Western Cape. Two clades were observed , which consisted of the first (LP, NW, GP, KZN and WC) and the second (MP and EC) clade isolates. Figure 13a represent the neighbor-joining tree of the isolates from MP and EC formed a cluster which was isolated from LP, NW, GP, KZN and WC with well supported bootstrap values of 99%. However, it shared high genetic simi larities with both clades with a high bootstrap support value of 99%. A maximum likelihood phylogenetic tree revealed strong bootstrap of 99% for a clade containing isolates of A. marginale msp4 gene from LP, NW, GP, KZN and WC. A strong bootstrap support of 82% from MP and EC isolates of South Africa (Figure 13b). These results were also observed in phylogenetic tree of A. marginale msp4 gene sequences from first clade (LP, NW, GP, KZN and WC) and the second clade (MP and EC) isolates of South Africa, North Africa, North and South America, Europe, Australia as well as Asia. 90 These phylogenetic trees were constructed using neighbor-joining and maximum- likelihood method (Figure 13a and 13b). Sequence comparison of the msp4 gene was recognized as one of the most powerful and precise methods for determining the phylogenetic relationships of A. marginale (de la Fuente et al., 2007a). South African strain had 100% nucleotide identity with isolates from Kenya, Zimbabwe and Australian. Even though they had highly shared the identity, South African strains remained in separate clusters. These strains shared 99% identity with isolates from North and South America, Europe and Asia (Appendix 4). Good representation of southern and northern hemisphere strains was observed, which further demonstrated that msp4 gene may serve as good marker for phylogeographic analysis (Figure 13a and 13b). In addition, these result suggest that South African cattle could be be infected with the same strains that infect cattle from West, East and North Africa, North and South America, Europe, Asia as well as Australia. The DNA sequences of msp4 gene revealed phylogeographical segregation with South African isolates and other isolates submitted in the NCBI database. MSP4 gene sequences provided phylogenetic resolution for intraspecific relationships among isolates. South African isolates demonstrated apparent phylogeograph ic information. The present study revealed that msp4 nucleotide sequences are sufficiently variable and like in other studies could be used to detect phylogeographic patterns broadly (de la Fuente et al., 2010). Therefore, the msp4 gene may be used as a marker for understanding the phylogeographic patterns and phylogenetic relationships of A. marginale in South Africa. 91 CHAPTER 6 GENERAL DISCUSSION, CONCLUSIONS AND RECOMMENDATIONS 6.1 General discussion The prevalence of bovine anaplasmosis in the various South African regions is often complex and may involve the bovine population immunity to A. marginale, the means of transmission (mechanical, biological by ticks, transplacental or a combination thereof) , as well as the genetic variability of A. marginale strains within a given area. In this study, data was not collected on the immunity of cattle populations for A. marginale. Certainly, tick transmission plays a very important role in the prevalence of A. marginale, and the distribution of tick species in South Africa has been determined by de Waal, (2000). The provinces with 100% prevalence of A. marginale infected cattle did not correlate with provinces with the highest or lowest A. marginale genetic diversity, but instead represented a range of diversity (Figure 7). It has been hypothesized that increased genetic variability may have an impact on A. marginale strains transmission (Palmer and Bryton, 2013). Alternately, A. marginale strains with low genetically variability may be lost after immune clearance. Interestingly, in the Western Cape the prevalence of A. margina/e was high while the genetic diversity was low, which could be due to low immunity of cattle population in the particular sample points. Overall , the prevalence of A. marginale in cattle from South Africa is high when compared with other regions of the world. In Brazil a recent study found prevalence of A. marginale in a cattle herd to be 70% (Pohl et al., 2013). Another study by de la Fuente et al. (2005a) reported the prevalence of A. marginale in cattle herds from Italy to range from 25 - 100%. 92 In these studies the prevalence and genetic diversity of A marginale were found to have a polynomial, and nonlinear correlation rather than a linear one. Subsequently, this finding most likely reflects the complexity of factors that may be influencing the prevalence of A marginale in the various regions of South Africa as evident in the province of Limpopo, North West and KwaZulu-Natal. This polynomial correlation suggests that increased genetic diversity impacted on the transmissibility of A marginale strains. The other major factor that could influence A margina/e prevalence in a given region is the immunity of a population . This factor is also related to genetic diversity because, as shown previously, a mechanism of infection exclusion may prevent cattle infected with one genotype from becoming infected with a second closely related one. However, widely separated genotypes apparently can be maintained in cattle as co-infections which could contribute to wider genetic diversity at the population level and impact on the prevalence of bovine anaplasmosis. Anap/asma marginale with low genetic variabil ity will likely become highly prevalent in a cattle population with predominately na"ive animals, without immunity as evidenced in the cattle population in the Western Cape. 6.2 Conclusions The regional prevalence of bovine anaplasmosis appears to be a result of complex factors, including the immunity of cattle populations to A marginale, the mode of transmission (biological by ticks, mechanical and/or transplacental) and the genetic variability of A. marginale strains within cattle populations. These factors have contributed to the challenges of developing effective diagnostic and control measures for A. marginale. Molecular tools are required for evaluating the genetic diversity of A. marginale. Developing an understanding of the molecular mechanisms underlying the generation of genetic diversity is crucial in implementation of effective control measures that are developed in concern with evolutionary changes of the pathogen which do not stay one step behind in pathogen evolution. 93 The first step in this molecular study of bovine anaplasmosis was to confirm that A. marginale was widespread in various regions in South Africa. The genetic diversity of A. marginale msp1a was then characterized , and found to arise from the evolution of extent tandem repeats, which under positive and negative selection, diversified as noted by the presence of newly reported tandem repeats variants. These new variants were most likely constricted by two forces: (1) the immune system and (2) biological transmission of A. marginale by ticks. Common msp1a tandem repeats were found among A. marginale strains in South Africa and other regions of the world which , in combination with the use of low variable msp 1a peptides, will likely be useful in the development of msp 1a based vaccine for more effective regional control of anaplasmosis. Gene sequences of msp4 provided sufficient variation to allow for the characterization of phylogeographic patterns on a broad scale. The South African clade of A. marginale msp4 gene sequences included isolates from West, East and North Africa, North and South America, Europe, Australia and Asia (Appendix 4). Finally the detection of diverse clades of A. marginale representing different geographic regions including the results that were obtained based on msp1a gene sequences may provide vital information for developing new novel vaccine approaches for control of anaplasmosis. This work presents information on the prevalence, genetic diversity and phylogenetic analysis of A. marginale in cattle throughout South Africa. Data compiled in this study will aid in the future management of A. marginale infection and may provide a molecular basis for targeted vaccine development. The information gained from this research provides an understanding of the epidemiology of A. margina/e in South Africa which will be fundamental toward the development of diagnostic assays and control measures for bovine anaplasmosis. 94 6.3 Recommendations 6.3.1 To further determine the species, distribution and population densities of the known tick vectors of A. marginale in the provinces of South Africa , and to correlate these tick populations with the prevalence of bovine anaplasmosis. 6.3.2 To characterize A. marginale strains in cattle populations in order to determine the strain diversity, and to conduct phylogenetic studies of these strains to gain insight into the origin and evolution of A. marginale in South Africa. 6.3.3 To identify the tick species infesting cattle in anaplasmosis endemic regions , and to determine the immunity of cattle in these areas for A. marginale infections. 6.3.4 To evaluate genes and gene products of the A. marginale strains, and to determine whether the diversity of msp4 gene is related to selective pressures. 95 REFERENCES AGNES, J. A. , BRAYTON, K. A. , LAFOLLETT, M., NORIMINE, J., BROWN, W . C. and PALMER, G.H. (2011 ). Identification of Anaplasma marginale outer membrane protein antigens conserved between sensu stricto strains and the live Anaplasma marginale ss. centrale vaccine. Infection and Immunity. 79: 1311-1318. AL-ADHAMI, B., SCANDRETT, W. W . B. , LOVANOV, V. and GAJADHAR, A. A. (2011 ). Serological cross reactivity between Anaplasma marginale and Ehrlichia species in naturally and experimentally infected cattle. Journal of Veterinary Diagnostic Investigation. 23: 1181-1188. ALLEMAN , A. R. and BARBET, A. F. (1996). Evaluation of Anaplasma marginale major surface protein 3 (MSP3) as a diagnostic test antigen. Journal of Clinical Microbiology. 34: 270-276. ALLEMAN , A. R. , PALMER, G. H., MCGUIRE, T. C., MCELWAIN, T. F. , PERRYMAN, L. E. and BARBET, A. F. (1997). Anaplasma marginale major surface protein 3 is encoded by a polymorphic, multigene family. Infection and Immunity. 65:1 56-163. ALLRED, D. R. , MCGUIRE, T. C., PALMER, G. H., LEIB, S. R. , HARKINS, T. M., MCELWAIN, T. F. and BARBET, A. F. (1990). Molecular basis for surface antigen size polymorphisms and conservation of a neutralization-sensitive epitope in Anaplasma marginale. Proceedings of the National Academy of Science USA. 87:3220-3224. 96 ALMAzAN , C., MEDRANO, C., ORTIZ, M. and DE LA FUENTE, J. (2008). Genetic diversity of Anaplasma marginale strains from an outbreak of bovine anaplasmosis in an endemic area. Veterinary Parasitology. 158: 103-109. AUBRY, P. and GEALE, D. W. (2011 ). A review of bovine anaplasmosis. Transboundary and Emerging Diseases. 58: 1-30. BAKER, M. K. , DUCASSE, F. B. W., SUTHERST, R. W. and MAYWALD, G. F. (1989). The seasonal tick populations on traditional and commercial cattle grazed at four altitudes in Natal. Journal of South African Veterinary Association. 60: 95-101 . BARBET, A. F., PALMER, G. H., MYLER, P. J. and MCGUIRE, T. C. (1987). Characterization of an immunoprotective protein complex of Anaplasma marginale by cloning and expression of the gene coding for polypeptide AM 105L. Infection and Immunity. 55: 2428-2435. BARBET, A. F., LUNDGREN, A. , JOOYOUNG, Y. I. , RURANGIRWA, F. R. and PALMER, G. H. (2000). Antigenic variation of Anaplasma marginale by expression of MSP2 mosaics. Infection and Immunity. 68: 6133-6138. BARBET, A. F., JOOYOUNG, Y. I., LUNDGREN, A. , MCEWEN, B. R. , BLOUIN, E. F. and KOCAN, K. M. (2001 ). Antigenic variation of Anaplasma marginale: major surface protein 2 diversity during cyclic transmission between ticks and cattle. Infection and Immunity. 69: 3057-3066. 97 BOCK, R. E. and DE VOS, A. J. (2001 ). Immunity following use of Australian tick fever vaccine: a review of the evidence. Australian Journal of Veterinary. 79: 832-839. BOWIE, M. V ., DE LA FUENTE, J., KOCAN, K. M., BLOUIN, E. F. and BARBET, A. F. (2002). Conservation of major surface protein 1 genes of Anaplasma marginale during cyclic transmission between ticks and cattle. Gene. 282: 95- 102. BRADWAY D. S., TORIONI DE ECHAIDE, S., KNOWLES, D. P., HENNAGER, S. G. and MCELWAIN T. F. (2001 ). Sensitivity and specificity of the complement fixation test for detection of cattle persistently infected with Anaplasma marginale. Journal of Veterinary Diagnostic Investigation. 13, 79-81 . BRAYTON, K. A., KNOWLES, D. P., MCGUIRE, T. C. and PALMER, G. H. (2001 ). Efficient use of a small genome to generate antigenic diversity in tick-borne ehrlichial pathogens. Proceedings of the National Academy of Science. 98: 4130-4135. BRAYTON, K. A. , KAPPMEYER, L. S., HERNDON, D. R., DARK, M. J., TIBBALS, D. L. , PALMER, G. H., MCGUIRE, T. C. and KNOWLES JR., D. P. (2005). Complete genome sequencing of Anaplasma marginale reveals that the surface is skewed to two superfamilies of outer membrane proteins. Proceedings of the National Academy of Science. 102: 844-849. BROCK, W . E., KLIEWER, I. 0 . and PEARSON, C. C. (1965). A vaccine for anaplasmosis. Journal of the American Veterinary Medical Association. 147: 948-951 . 98 BROWN, W. C., ZHU, D., SHKAP, V., MCGUIRE, T. C., BLOUIN, E. F., KOCAN, K. M. and PALMER, G. H. (1998). The repertoire of Anaplasma marginale antigens recognized by CD4p T-lymphocyte clones from protectively immunized cattle is diverse and includes major surface protein 2 (MSP-2) and MSP-3. Infection and Immunity. 66: 5414-5422. BROWN, W. C., MCGUIRE, T. C., ZHU, D., LEWIN, H. A. , SOSNOW, J. and PALMER, G. H. (2001 ). Highly conserved regions of the immunodominant major surface protein 2 of the gene group II ehrlichial pathogen Anaplasma marginale are rich in naturally derived CD4+ T lymphocyte epitopes that elicit strong recall responses. Journal of Immunology. 166: 1114-1124. BROWN, W . C., BRAYTON, K. A. , STYER, C. M. and PALMER, G. H. (2003). The hypervariable region of Anaplasma marginale major surface protein 2 (MSP2) contains multiple immunodominant CD4+ T lymphocyte epitopes that elicit variant-specific proliferative and IFN-y responses in MSP2 vaccinates. Journal of Immunology. 170: 3790-3798. BRYSON, N. R. , TICE, G. A. , HORAK, I. G., STEWART, C. G. and DU PLESSIS, B. J. A. (2002). lxodid ticks on cattle belonging to small-scale farmers at 4 communal grazing areas in South Africa. Journal of the South African Veterinary Association. 73: 98-103. BUCHAN, D. W ., WARD, S. M., LOBLEY, A. E., NUGENT, T. C., BRYSON, K. and JONES, D. T. (2010). Protein annotation and modelling servers at University College London. Nucleic Acids Research. 38: W563-568. 99 CABEZAS-CRUZ, A. , PASSOS, L. M. F., LIS, K. , KENNEil , R. , VALDES, J. J., FERROLHO, J., TONK, M., POHL, A. E. , GRUBHOFFER, L. , ZWEYGARTH, E., SHKAP, V ., RIBEIRO, M. F. B., ESTRADA-PENA, A. , KOCAN , K. M. and DE LA FUENTE, J. (2013). Functional and Immunological Relevance of Anaplasma marginale Major Surface Protein 1 a Sequence and Structural Analysis. PloS One. 8: 1-13. CAMACHO NUEZ, M., MONOZ, M. L. , SUAREZ, C. E., MCGUIRE, T. C., BROWN, W. C. and PALMER, G. H. (2000). Expression of polymorphic msp1 a genes during acute Anaplasma marginale rickettsemia. Infection and Immunity. 68: 1946-1952. CARELLI , G., DECARO, N., LORUSSO, A. , ELIA, G., LORUSSO, E., MARI , V ., CECI , L. and BUONAVOGLIA, C. (2007). Detection and quantification of Anaplasma marginale DNA in blood samples of cattle by real-time PCR. Veterinary Microbiology. 124: 107- 114. CARRENO, A. D., ALLEMAN, A. R. , BARBET, A. F., PALMER, G. H., NOH, S. M. and JOHNSON, C. M. (2007). In vivo endothelial cell infection by Anaplasma marginale. Veterinary Pathology. 44: 116-118. CHOMEL, B. B., CARNICIU, M. L. , KASTEN, R. W ., CASTELLI , P. M., WORK, T. M. and JESSUP, D. A. (1994). Antibody prevalence of eight ruminant infectious diseases in California mule and black-tailed deer (Odocoileus hemionus). Journal of Wildlife Diseases. 30: 51-59. 100 COETZEE, J. F., APLEY, M. D., KOCAN, K. M., RURANGIRWA, F. R. , and VAN DONKERSGOED, J. (2005). Comparison of three oxytetracycline regimes for the treatment of persistent Anaplasma marginale infections in beef cattle. Veterinary Parasitology. 127: 61-73. COETZEE, J. F., APLEY, M. D., and KOCAN, K. M. (2006). Comparison of the efficacy of enrofloxacin, imidocarb, and oxytetracycline for clearance of persistent. Anaplasma marginale infections in cattle. Journal of Veterinary Therapeutics. 7: 347- 360. COETZEE, J. F., SCHMIDT, P. L., APLEY, M. D., REINBOLD, J. B. and KOCAN, K. M. (2007). Comparison of the complement fixation test and competitive ELISA for serodiagnosis of Anaplasma marginale infection in experimentally infected steers. American Journal of Veterinary Research. 68: 872-878. COETZER, J. A. W ., THOMSON, G. R. and TUSTIN, R. C. (1994). Infectious Diseases of Livestock with Special Reference to Southern Africa. Oxford University Press, Cape Town. COLEMAN, P. G., PERRY, B. D. and WOOLHOUSE, M. E. J. (2001). Endemic stability-a veterinary idea applied to human public health . Lancet. 357: 1284-1286. D'HAESE, L. , PENNE, K. and ELYN, R. (1999). Economics of theileriosis control in Zambia. Tropical Medicine and International Health. 4: 49-57. 101 DARK, M. J., AL-KHEDERY, B. and BARBET, A. F. (2011 ) Multistrain genome analysis identifies candidate vaccine antigens of Anaplasma marginale. Vaccine. 29: 4923- 4932. DE CASTRO J. J. (1997). Sustainable tick and tick-borne disease control in livestock improvement in developing countries. Veterinary Parasitology. 77: 77-97. DE LA FUENTE, J., GARCIA-GARCIA, J. C., BLOUIN, E. F., RODRIGUEZ, S. D., GARCIA, M. A. , and KOCAN , K. M. (2001 a). Evolution and function of tandem repeats in the major surface protein 1 a of the ehrlichial pathogen Anaplasma marginale. Animal Health Research and Review. 2: 163-173. DE LA FUENTE, J., GARCIA-GARCIA, J. C., BLOUIN, E. F., and KOCAN, K. M. (2001 b). Differentia l adhesin of major surface proteins 1a and 1 b of the ehrlichial cattle pathogen Anaplasma marginale to bovine erythrocytes and tick cells . International Journal of Parasitology. 31: 145-153. DE LA FUENTE, J., GARCIA-GARCIA, J. C., BLOUIN, E. F., and KOCAN, K. M. (2001 c) . Major surface protein 1 a effects tick infection and transmission of the ehrlichial pathogen Anaplasma marginale. International Journal of Parasitology. 31: 1705-1714. DE LA FUENTE, J., VAN DEN BUSSCHE, R. A. , and KOCAN , K. M. (2001d). Molecular phylogeny and biogeography of North American strains of Anaplasma marginale (Rickettsiaceae: Ehrlichieae). Veterinary Parasitology. 97: 65-76. 102 DE LA FUENTE, J., VAN DEN BUSSCHE, R. A. , GARCIA-GARCIA, J. C., RODRIGUEZ, S. D., GARCIA, M. A. , GUGLIELMONE, A. A. , MANGOLD, A. J., FRICHE PASSOS, L. M ., BLOUIN, E. F. and KOCAN , K. M. (2002a). Phylogeography of New World st rains of Anaplasma marginale (Rickettsiaceae: Ehrlichieae) based on major surface protein sequences. Veterinary Microbiology. 88: 275-285. DE LA FUENTE, J., GARCIA-GARCIA, J. C., BLOUI N, E. F., SALIKI , J. T., and KOCAN, K. M. (2002b). Infection of tick cells and bovine erythrocytes with one genotype of the intrace llular ehrl ichia Anaplasma marginale excludes infection with other genotypes. Clinical and Diagnostic Laboratory Immunology. 9: 658-668. DE LA FUENTE, J., GARCIA-GARCIA, J. C., BLOUIN, E. F. and KOCAN , K. M. (2003a). Characterization of the functional domain of major surface protein 1 a involved in adhesin of the rickettsia Anaplasma marginale to host cells. Veterinary Microbiology. 91: 265-283. DE LA FUENTE, J., VAN DEN BUSSCHE, R. A. , PRADO, T . M. and KOCAN, K. M. (2003b). Anaplasma marginale msp1a genotypes evolved under positive selection pressure but are not markers for geographic isolates. Journal of Clinical Microbiology. 41: 1609-1616. DE LA FUENTE, J., LEW, A. , LUTZ, H., MELI, M. L. , HOFMANN-LEHMANN, R. , SHKAP, V ., MOLAD, T. , MANGOLD, A. J., ALMAZAN, C., NARANJO, V., GORTAZAR, C., TORINA, A. , CARACAPPA, S., GARCIA-PEREZ, A. L., BARRAL, M., OPORTO, B., CECI , L., CARELLI , G ., BLOUIN, E. F. and KOCAN, K. M. (2005a). Genetic diversity of Anaplasma species major surface proteins and implications for anaplasmosis serodiagnosis and vaccine development. Animal Health Research and Review. 6: 75- 89. 103 DE LA FUENTE, J., TORINA, A. , CARACAPPA, S., TUMINO, G., FURLA, R. , ALMAZAN, C. and KOCAN, K. M. (2005b). Serologic and molecular characterization of Anaplasma species infection in farm animals and ticks from Sicily. Veterinary Parasitology. 133: 357-362. DE LA FUENTE, J., MASSUNG, R. B., WONG, S. J., CHU, F. K., LUTZ, H., MELI , M., VON LOEWENICH, F. D., GRZESZCZUK, A. , TORINA, A. , CARACAPPA, S., MANGOLD, A. J., NARANJO, V., STUEN, S. and KOCAN, K. M. (2005c). Sequence analysis of the msp4 gene of Anaplasma phagocytophilum strains. Journal of Clinical Microbiology. 43: 1309-1 317. DE LA FUENTE, J., NARANJO, V., RUIZ-FONS, F., HOFLE, U., FERNANDEZ DE MERA, I. G., VILLANUA, D., ALMAZAN, C., TORINA, A. , CARACAPPA, S., KOCAN, K. M. and GORTA ZAR, C. (2005d). Potential vertebrate reservoir hosts and invertebrate vectors of Anaplasma marginale and A. phagocytophilum in central Spain. Vector-Borne Zoonosis and Diseases. 5: 390-401 . DE LA FUENTE, J., RUYBAL, P., MTSHALI , M. S., NARANJO, LI SHUQING, L. , MANGOLD, A. J., RODRIGUEZ, S. D., JIMENEZ, R. , VICENTE, J., MORETTA, R. , TORINA, A. , ALMAZAN, C. , MBATI, P. M., FARBER, M., GORTAZAR, C. and KOCAN, K. M. (2007a). Analysis of world strains of Anaplasma marginale using major surface protein 1a repeat sequences. Veterinary Microbiology. 119: 382-390. DE LA FUENTE, J., ATKINSON, M. W., NARANJO, V., FERNANDEZ DE MERA, I. G., MANGOLD, A. J., KEATING, K. A. and KOCAN, K. M. (2007b). Sequence analysis of the msp4 gene of Anaplasma ovis strains. Veterinary Microbiology.119: 375-81 . 104 DE LA FUENTE, J., KOCAN, K. M., BLOUIN, E. F., ZIVKOVIC, Z., NARANJO, V., ALMAZAN, C., ESTEVES, E., JONGEJAN, F., DAFFRE, S. and MANGOLD, A. J. (2010). Functional genomics and evolution of tick- Anaplasma interactions and vaccine development. Veterinary Parasitology. 167: 175- 186. DE WAAL, D. T. (2000). Anaplasmosis control and diagnosis in South Africa. Annals of the New York Academy of Science. 916: 474-483. DECARO, N., ELIA, G., DESARIO, C., ROPERTO, S., MARTELLA, V., CAMPOLO, M., LORUSSO, A. , CAVALLI , A. and BUONAVOGLIA, C. (2006a). A minor groove binder probe real-time PCR assay for discrimination between type 2-based vaccines and field strains of canine parvovirus. Journal of Virology Methods. 136: 65-70. DECARO, N. , ELIA, G., MARTELLA, V., CAMPOLO, M., DESARIO, C., CAMERO, M., CIRONE, F., LORUSSO, E., LUCENTE, M.S., NARCISI , D., SCALIA, P. and BUONAVOGLIA, C. (2006b). Characterisation of the canine parvovirus type 2 variants using minor groove binder probe technology. Journal of Virology Methods. 133: 92-99. DECARO, N., MARTELLA, V., ELIA, G., DESARIO, C., CAMPOLO, M., BUONAVOGLIA, D., BELLACICCO, AL., TEMPESTA, M. and BUONAVOGLIA, C. (2006c). Diagnostic tools based on minor groove binder technology for rapid identification of vaccine and field strains of canine parvovirus type 2b. Journal of Virology Methods. 138: 10--16. 105 DECARO, N., CARELLI , G., LORUSSO, E., LUCENTE, M.S., GRECO G., LORUSSO, A. , RADOGNA, A. , CECI , L. and BUONAVOGLIA, C. (2008). Duplex real-time polymerase chain reaction for simultaneous detection and quantification of Anaplasma marginale and Anaplasma centrale. Journal of Veterinary Diagnostic Investigation. 20: 606-611. DELPORT, W ., POON, A. F., FROST, S. D. and KOSAKOVSKY POND, S. L. (2010). Datamonkey 2010: suite of phylogenetic analysis tools for evolutionary biology. Bioinformatics. 26: 2455-2457. DOLD, A. P. and COCKS, M. L. (2001 ). Traditional veterinary medicine in the Alice district of the Eastern Cape Province, South Africa. South African Journal of Science. 97: 375-379. DREHER, U. M., DE LA FUENTE, J., HOFMANN-LEHMANN, R. , MELI, M. K. , PUSTERIA, N., KOCAN, K. M., WOLDEHIWET, A. , REGULA, G. and STAERK, K. D. C. (2005). Serologic cross reactivity between Anaplasma marginale and Anaplasma phagocytophilum. Clinical Vaccine Immunology. 12: 1177-1183. DREYER, K., FOURIE, L. J., and KOK, D. J. (1998). Epidemiology of tick-borne diseases of cattle in Botshabelo and Thaba Nchu in the Free State Province. Onderstepoort Journal of Veterinary Research. 65: 285-289. 106 DU PLESSIS, J. L., DE WAAL, D. T. and STOLTSZ, W . H. (1994). A survey of the incidence and importance of the tick-borne diseases heartwater, redwater and anaplasmosis in the heartwater -endemic regions of South Africa. Onderstepoort Journal of Veterinary Research. 61: 295-301 . DUMLER, J., BARBET, A. , BEKKER, C., DASCH, G., PALMER, G., RAY, S., RIKIHISA, Y. and RURANGIRWA, F. (2001 ). Reorganization of genera in the families Rickettsiaceae and Anaplasmataceae in the order Rickettsiales: unification of some species of Ehrlichia with Anaplasma, Cowdria with Ehrlichia and Ehrlichia with Neorickettsia, descriptions of six new species combinations and designation of Ehrlichia equi and 'HE agent' as subjective synonyms of Ehrlichia phagocytophila. International Journal of Systematics and Evolution Microbiology. 51: 2145-2165. DUNNING HOTOPP, J. C., LIN, M., MADUPU, R. , CRABTREE, J., ANGIUOLI, S. V ., EISEN, J., SESHADRI, R. , REN, Q., WU, M., UTTERBACK, T. R. , SMITH, S., LEWIS, M., KHOURI , H., ZHANG, C., NIU, H., LIN, Q., OHASHI, N., ZHI , N., NELSON, W ., BRINKAC, L. M., DODSON, R. J., ROSOVITZ, M. J., SUNDARAM, J., DAUGHERTY, S. C., DAVIDSEN, T. , DURKIN, A. S., GWINN, M., HAFT, D. H., SELENGUT, J. D., SULLIVAN, S. A. , ZAFAR, N., ZHOU, L., BENAHMED, F., FORBERGER, H., HALPIN, R. , MULLIGAN , S., ROBINSON, J., WHITE, 0 ., RIKIHISA, Y. and TETTELIN, H. (2006). Comparative genomics of emerging human ehrlichiosis agents. PLoS Genetics. 2 (2): e21 . ELIA, G., SAVINI, G., DECARO, N., MARTELLA, V., TEODORI, L., CASACCIA, C., DI GIALLEONARDO, L. , LORUSSO, E., CAPORALE, V. AND BUONAVOGLIA, C. (2008). Use of real-time RT-PCR as a rapid molecular approach for differentiation of field and vaccine strains of bluetongue virus serotypes 2 and 9. Molecular Cellular and Probes. 22: 38-46. 107 ELLERY, W. N., SCHOLES, R. J. and SCHOLES, M. C. (1995). The distribution of sweetveld and sourveld in South Africa's grassland biome in relation to environmental factors. African Journal of Range and Forage Science. 12: 38-45. ERIKS, I. S., STILLER, D. and PALMER, G.H. (1993). Impact of persistent Anaplasma marginale rickettsemia on tick infection and transmission . Journal of Clinical Microbiology. 31: 2091 . ESTRADA-PENA, A , NARANJO, V., ACEVEDO-WHITEHOUSE, K., MANGOLD, A J., KOCAN, K. M. and DE LA FUENTE, J. (2009). Phylogeographic analysis reveals association of tick-borne pathogen, Anaplasma marginale, Msp1a sequences with ecological traits affecting tick vector performance. BioMed Central Biology. 57: 1-13. ExPaSy Translation Tool 2011 . Available from: http://expasy.hcuge.ch/ tools/dna.html. FELSENSTEIN, J. (2004). Inferring Phylogenies. Sunderland Massachusetts: Sinauer Associates. FOURIE, L. J. and HORAK, I. G. (1991). The seasonal activity of adult ixodid ticks on Angora goats in the south western Orange Free State. Journal of the South African Veterinary Association. 62: 104-106. FOURIE, L. J., KOK, D. J. and HEYNE, H. (1996). Adult ixodid ticks on two cattle breeds in the southwestern Free State, and their seasonal dynamics. Onderstepoort Journal of Veterinary Research. 63: 19-23. 108 FRENCH, D. F., MCELWAIN, T. F., MCGUIRE, T. C. and PALMER, G. H. (1998). Expression of Anaplasma marginale Major Surface Protein 2 variants during persistent cyclic rickettsemia. Infection and Immunity. 66: 1200-1207. FRENCH, D. M., BROWN, W. C. and PALMER, G. H. (1999). Emergence of Anaplasma marginale antigenic variants during persistent rickettsemia. Infection and Immunity. 67: 5834-5840. FUTSE, J. E., UETI, M. W., KNOWLES, D. P., JR. and PALMER, G. H. (2003). Transmission of Anaplasma marginale by Boophilus microplus: retention of vector competence in the absence of vector-pathogen interaction. Journal of Clinical Microbiology. 41: 3829-3834. GARCIA-GARCIA, J . C., DE LA FUENTE, J ., BLOUIN, E. F., HALBUR, T. , ONET, V. C., SALIKI , J. T. and KOCAN, K. M. (2004a). Differential expression of the msp 1a gene of Anaplasma marginale occurs in bovine erythrocytes and tick cells. Veterinary Microbiology. 98: 261-272. GARCIA-GARCIA, J. C., DE LA FUENTE, J., KOCAN, K. M., BLOUIN, E. F., HALBUR, T., ONET, V. C. and SALIKI, J. T. (2004b). Mapping of B-cell epitopes in the N-terminal repeated peptides of Anaplasma marginale major surface protein 1a and characterization of the humeral immune response of cattle immunized with recombinant and whole organism antigens. Veterinary Immunology and lmmunopathology. 98: 137- 151 . 109 GARCIA-BORONAT, M., DIEZ-RIVERO, C. M., REINHERZ, E. L. , and RECHE, P. A. (2008). PVS: a web server for protein sequence variability analysis tuned to facilitate conserved epitope discovery. Nucleic Acids Research. 36: 35-41 . GATES, N. I. and WESCOTT, R. B. (2000). Parasites of Cattle. WSU, cru84.cahe.wsu.edu/cgibin/pubs/EB1742.html, Accessed 14 August 2013. GE, N., KOCAN, K. M., BLOUIN, E. F. and MURPHY, G. L. (1996). Developmental studies of Anaplasma marginale (Rickettsiales: Anaplasmataceae) in male Dermacentor andersoni (Acari : lxodidae) infected as adults by using nonradioactive in situ hybridization and microscopy. Journal of Medical Entomology. 33: 911-920. HALL, B. G. (2011 ). Phylogenetic Trees Made Easy: A How-To Manual. 4th Edition. Sinauer, Assoc. Sunderland, MA. HALL, B. G. (2013). Building Phylogenetic Trees from Molecular Data with MEGA. Molecular Biology Evolution . 30: 1229-1235. HALL, B. G. and BARLOW, M. (2006). Phylogenetic Analysis as a Tool for Molecular Epidemiology of Infectious Diseases. Annals of Epidemiology. 16: 157-169. HALL, B. G., PIKIS, A. and THOMPSON, J. (2009). Evolution and biochemistry of Family 4 glycosidases: implications for assigning enzyme function in sequence annotations. Molecular Biology Evolution. 26: 2487-2497. 110 HAMOU, S. A. , RAHALi , T. , SAHIBI, H., BELGHYTI , D., LOSSON, 8 ., GOFF, W. and RHALEM, A. (2012). Molecular and serological prevalence of Anaplasma marginale in cattle of North Central Morocco. Research in Veterinary Science. 93: 1318-1323. HART, L. T. , LARSON, A. D., DECKER, J. L. , WEEKS, J. P., and CLANCY, P. L. (1981) . Preparation of intact Anaplasma marginale devoid of host cell antigen. Current Microbiology. 5: 95-100. HOFMANN-LEHMANN, R. , MELI , M. L. , DREHER, U. M., GONCZI, E. , DEPLAZES, P., BRAUN, U., ENGELS, M., SCHUPBACH, J., JORGER, K. , THOMA, R. , GRIOT, C., STARK, K. D. C., WILLI , B., SCHMIDT, J., KOCAN, K. M. and LUTZ H. (2004). Concurrent infections with vector-borne pathogens associated with fatal haemolytic anemia in a cattle herd in Switzerland. Journal of Clinical Microbiology. 42: 3775-3780. HORAK, I. G., KNIGHT, M. M. and WILLIAMS, E. J. (1991 ). Parasites of domestic and wild animals in South Africa. XXVlll. Helminth and arthropod parasites of Angora goats and kids in Valley Bushveld. Onderstepoort Journal of Veterinary Research. 58: 253- 260. HORAK, l.G. (1999). Parasites of domestic and wild animals in South Africa. XXXVll. lxodid ticks on cattle on Kikuyu grass pastures and in Valley Bushveld in the Eastern Cape Province. Onderstepoort Journal of Veterinary Research. 66: 175- 184. 111 HORNOK, S., FOLDVARI , G., ELEK, V ., NARANJO, V ., FARKAS, R. , and DE LA FUENTE, J. (2008). Molecular identification of Anaplasma marginale and rickettsial endosymbionts in blood-sucking flies (Diptera: Tabanidae, Muscidae) and hard ticks (Acari : lxodidae). Veterinary Parasitology. 154: 354-359. HOWDEN, K. J., and GEALE, D. W. (2010). An update on bovine anaplasmosis (Anaplasma marginale) in Canada. Journal of Canada Veterinary Research. 51: 837- 840. INOKUMA, H., YUTAKA, T. , KAMIO, T., RAOULT, D. and BROUQUI , P. (2001). Analysis of the 16S rRNA gene sequence of Anaplasma centrale and its phylogenetic relatedness to other Ehrlichiae. Clinical and Diagnostic Laboratory Immunology. 8: 241- 244. JONSSON, N. N. and REID, S. W . J. (2000). Global climate change and vector borne diseases. Guest editorial. Veterinary Journal. 160: 87-89. JONSSON, N. N., DAVIS, R. and DE WITT, M. (2001). An estimate of the economic effects of cattle tick (Boophilus microplus) infestation on Queensland dairy farms. Australian Veterinary Journal. 79: 826-831. JONSSON, N. N., BOCK, R. E. , JORGENSEN, W . K., MORTON, J. M. and STEAR, M. J. (2012). Is endemic stability of tick-borne disease in cattle a useful concept? Trends in Parasitology. 28: 85-89. 112 KANO, F. S., VIDOTTO, 0 ., PACHECO, R. C. and VIDOTTO, M. C. (2002). Antigenic characterization of Anaplasma marginale isolates from different regions of Brazil. Veterinary Microbiology. 87: 131-138. KAUFMAN, P. E., KOEHLER, P. G. and BUTLER, J. F. (2006). External Parasites on Beef Cattle. Entomology and Nematology Department, Florida Cooperative Extension Service, Institute of Food and Agricultural Sciences, University of Florida, http://edis.ifas.ufl.edu/IG130, Accessed 12 May 2013. KIESER, S. T., ERIKS, I. S. and PALMER, G. H. (1990). Cyclic rickettsemia during persistent Anaplasma marginale infection of cattle. Infection and Immunity. 58: 1117- 1119. KIMURA, M. (1980). A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. Journal of Molecular Evolution. 16: 111-120. KNOWLES, D., TORIONI DE ECHAIDE, S., PALMER, G. , MCGUIRE, T. , STILLER, D. and MCELWAIN, T. (1996). Antibody against an Anaplasma marginale MSP5 epitope common to tick and erythrocyte stages identifies persistently infected cattle. Journal of Clinical Microbiology. 34: 2225-2230. 113 KOCAN, K. M., GOFF, W. L., STILLER, D., CLAYPOOL, P. L. , EDWARDS, W., EWING, S. A. , HAIR, J. A. and BARRON, S. J. (1992). Persistence of Anaplasma marginale (Rickettsiales: Anaplasmataceae) in male Oermacentor andersoni (Acari : lxodidae) transferred successively from infected to susceptible calves. Journal of Medical Entomology. 29: 657-668. Kocan, K.M., (1999). Parasitol. Today 15 (7) poster. KOCAN, K. M., BLOUIN, E. F. and BARBET, A. F. (2000). Anaplasmosis control : past, present and future. Annals of the New York Academy of Science. 916: 501-509. KOCAN, K. M., DE LA FUENTE, J., BLOUIN, E.F., and GARCIA-GARCIA, J. C. (2002). Adaptations of the tick-borne pathogen, Anaplasma marginale , for survival in cattle and ticks. Experimental and Applied Acarology. 28: 9-25. KOCAN, K. M., DE LA FUENTE, J., GUGLIELMONE, A. A. and MELENDEZ, R. D. (2003). Antigens and alternatives for control of Anaplasma marginale infection in cattle. Clinical Microbiology Review. 16: 698-712. KOCAN, K. M., DE LA FUENTE, J., BLOUIN, E. F. and GARCIA-GARCIA, J. C. (2004). Anaplasma marginale (Rickettsiales: Anaplasmataceae): recent advances in defining host pathogen adaptations of a tick-borne rickettsia . Parasitology. 129: S285- 300. KOCAN, K. M., DE LA FUENTE, J. and BLOUIN, E. F. (2008). Advances toward understanding the molecular biology of the Anap/asma-tick interface. Front Biosci Journal. 13: 7032-7045. 114 KOGAN , K. M., DE LA FUENTE, J., BLOUIN, E. F., COETZEE, J. F. and EWING, S. A. (2010). The natural history of Anaplasma marginale. Veterinary Parasitology. 167: 95- 107. KUTTLER, K.L. (1984). Anaplasma infections in wild and domestic ruminants: a review. Journal of Wildlife Diseases. 20: 12-20. LEW, A. E. , BOCK, R. E., MINCHIN, C. M. and MASAKA, S. (2002). A msp1a polymerase chain reaction assay for specific detection and differentiation of Anaplasma marginale isolates. Veterinary Microbiology. 86: 325-335. LOHR, C. V ., BRAYTON, K. A. , SHKAP, V., MOLDA, T. , BARBET, A. F., BROWN, W. C. and PALMER, G. H. (2002). Expression of Anaplasma marginale major surface protein 2 operon-associated proteins during mammalian and arthropod infection. Infection and Immunity. 70: 6005-6012. LOPEZ, J. E., SIEMS, W . F., PALMER, G. H., BRAYTON, K. A. , MCGUIRE, T . C., NORIMINE, J. and BROWN, W . C. (2005). Identification of novel antigenic proteins in a complex Anaplasma marginale outer membrane immunogen by mass spectrometry and genomic mapping. Infectious Immunology. 73: 8109-8118. LOPEZ, J. E., BEARE, P. A. , HEINZEN, R. A. , NORIMINE, J., LAHMERS, K. K. , PALMER, G. H. and BROWN, W . C. (2008). High-throughput identification of T- lymphocyte antigens from Anaplasma marginale expressed using in vitro transcription and translation. Journal of Immunological Methods. 332: 129-141 . 115 MACMILLAN, H., BRAYTON, K.A. , PALMER, G.H., MCGUIRE, T.C., MUNSKE, G., SIEMS, W.F. and BROWN, W.C. (2006). Analysis of the Anaplasma marginale major surface protein 1 complex protein composition by tandem mass spectrometry . Journal of Bacteriology. 188: 4983-4991 . MAPIYE C., CHIMONYO M., DZAMA K. , RAATS J. G. and MAPEKULA M. (2009). Opportunities for improving Nguni cattle production in the smallholder farming systems of South Africa. Livestock Science. 124: 196-204. MARTINEZ, M. L., MACHADO, M. A , NASCIMENTO, C. S., SILVA, M. V. G. B., TEODORO, R. L. , FURLONG, J., PRATA, M. C. A , CAMPOS, A L. , GUIMARAES, M. F. M., AZEVEDO, A L. S., PIRES, M. F. A and VERNEQUE, R. S. (2006). Association of BoLA-DRB3.2 alleles with tick (Boophilus microplus) resistance in cattle. Genetics and Molecular Research. 5: 513-524. MARUFU, M. C. (2008). Prevalence of ticks, tick-borne diseases in cattle on communal rangelands in the highland areas of the Eastern Cape Province. M.Sc., Thesis, University of Fort Hare, Alice, South Africa. MARUFU, M. C., CHIMONYO, M., DZAMA, K. and MAPIYE C. (2011). Tick prevalence in cattle raised on sweet and sour rangelands in semi-arid areas. Tropical Animal Health and Production. 43: 307-313. MARUFU, M. C. (2014). Mechan isms of resistance to Rhipicephalus ticks in Nguni cattle reared in the semiarid areas of South Africa. Doctoral Thesis, University of KwaZulu-Natal, South Africa. 11 6 MATTIOLI , R. C., PANDEY, V. S., MURRAY, M. and FITZPATRICK, J. L. (2000). Review: lmmunogenetic influences on tick resistance in African cattle with particular reference to trypanotolerant N'Dama (Bos taurus) and trypanosusceptible Gobra zebu (Bos indicus) cattle. Acta Tropica. 75: 263-277. MCCORKLE-SHIRLEY, S., HART, L. T. , LARSON, A. D., TODD, W . J. and MYHAND, J.D. (1985). High-yield preparation of purified Anaplasma marginale from infected bovine red blood cells. American Journal of Veterinary Research. 46: 1745-1747. MCGAREY, D. J. and ALLRED, D. R. (1994). Characterization of hemagglutinating components on the Anaplasma marginale initial body surface and identification of possible adhesins. Infection and Immunity. 62: 4587-4593. MCGAREY, D. J., BARBET, A. F., PALMER, G. H., MCGUIRE, T. C. and ALLRED, D. R. (1994 ). Putative adhesins of Anaplasma marginale: major surface polypeptides 1a and 1 b. Infection and Immunity. 62: 4594-4601 . MCGUIRE, T. C., PALMER, G. H., GOFF, W. L., JOHNSON, M. I. and DAVIS, W . C. (1984). Common and isolate restricted antigens of Anaplasma marginale detected with monoclonal antibodies. Infection and Immunity. 45: 697-700. MCGUIRE, T. C., DAVIS, W . C., BRASSFIELD, A. L. , MCELWAIN, T.F. and PALMER, G. H. (1991 ). Identification of Anaplasma marginale long-term carrier cattle by detection of serum antibody to isolated MSP-3. Journal of Clinical Microbiology. 29: 788-793. 117 MEEUS, P. F. M. and BARBET, A. F. (2001 ). Ingenious gene generation. Trends Microbiology. 9: 353-355. MINJAUW, B., PERRY, B. D., KRUSKA, R. , PETER, T. F. , CHAMBOKO, T. , MAHAN, S. M., MEDLEY, G. F. and O'CALLAGHAN, C. J. (1998). Economic impact assessment of heartwater in southern Africa . In : Proceedings of the 9th International Conference of the Association of Institutions of Tropical Veterinary Medicine (AITVM), Harare, Zimbabwe. MINJAUW, B. and MCLEOD, A. (2003). The impact of ticks and tick-borne diseases on the livelihood of small-scale and marginal livestock owners in India and eastern and southern Africa. Research report, DFID Animal Health Programme, Centre for Tropical Veterinary Medicine, University of Edinburgh, UK. Tick-borne diseases and poverty. MOLAD, T. , MAZUZ, M. L. , FLEIDEROVITZ, L. , FISH, L. , SAVITSKY, I. , KRIGEL, Y., LEIBOVITZ, B., MOLLOY, J., JONGEJAN, F. and SHKAP, V. (2006). Molecular and serological detection of A. centrale- and A. marginale-infected cattle grazing within an endemic area. Veterinary Microbiology. 113: 5~2. MOLAD, T., FLEIDROVICH, L., MAZUZ, M., FISH, L. , LEIBOVITZ, B., KRIGEL, Y., SH KAP, V. (2009). Genetic diversity of major surface protein 1a of Anaplasma marginale in beef cattle. Veterinary Microbiology. 136: 54-60. 118 MOLLOY, J. B. , BOWLES, P. M., KNOWLES, D. P., MCELWAIN, T. F., BOCK, R. E., KINGSTON, T. G., BLIGHT, G. W. and DALGLIESH, R. J. (1999). Comparison of a competitive inhibition ELISA and the card agglutination test for detection of antibodies to Anaplasma marginale and Anaplasma centrale in cattle. Australian Veterinary Journal. 77: 245-249. MTSHALI , M. S., DE WAAL, D. T. and MBATI, P. A (2004). A sero-epidemiological survey of blood parasites in cattle in the north-eastern Free State, South Africa. Onderstepoort Journal of Veterinary Research. 71: 67-75. MTSHALI, M. S., DE LA FUENTE, J., RUYBAL, P., KOCAN, K. M., VICENTE, J., MBATI, P. A , SHKAP, V., BLOUIN, E. F., MOHALE, N. E., MOLOI, T. P., SPICKETT, A M. and LATIF, A A (2007). Prevalence and genetic diversity of Anaplasma marginale strains in cattle in South Africa. Zoonotic Public Health. 54: 23-30. MTSHALI, M. S. and MTSHALI , P. S. (2013). Molecular diagnosis and phylogenetic analysis of Babesia bigemina and Babesia bovis hemoparasites from cattle in South Africa. BioMed Central Veterinary Research. 9: 154. MUCHENJE, V., DZAMA, K., CHIMONYO, M., RAATS J. G. and STRYDOM, P. E. (2008). Tick susceptibility and its effects on growth performance and carcass characteristics of Nguni, Bonsmara and Angus steers raised on natural pasture. Animal. 2: 298-304. 119 MUKHEBI, A W ., WATHANGA, J. , PERRY, B. D., IRVIN, A D. and MORZARIA, S. P. (1989). Financial analysis of East Coast fever control strategies on beef production under farm conditions. Veterinary Record. 125: 456-459. MUTSHEMBELE, AM., CABEZAS-CRUZ, A , MTSHALI , M.S., THEKISOE, O.M.M., GALINDO, R.C. AND DE LA FUENTE, J. (2014). Epidemiology and evolution of the genetic variability of Anaplasma marginale in South Africa. Ticks and Tick -borne diseases. 5: 24-631 . NARANJO, V., RUIZ-FONS, F. , HOFLE, U., FERNANDEZ DE, M. I. G., VILLANUA, D., ALMAZAN, C., TORINA, A , CARACAPPA, S., KOCAN, K. M., GORTAZAR, C. and DE LA FUENTE, J. (2006). Molecular epidemiology of human and bovine anaplasmosis in southern Europe. Annals of the New York Academy Science. 1078: 95- 99. NDOU, R. V., DIPHAHE, T. P., DZOMA, B. M. and MOTSEI , L. E. (2010). The seroprevalence and endemic stability of anaplasmosis in cattle around Mafikeng in the North West Province, South Africa. Veterinary Research. 3: 1-3. NEI , M. and KUMAR, S. (2000). Molecular evolution and phylogenetics. Oxford University Press, New York. Pg. 10-50. NELSON, C. M., HERRON, M. J., FELSHEIM, R. F., SCHLOEDER, B. R. , GRINDLE, S. M., CHAVEZ, A 0 ., KURTTI, T. J. and MUNDERLOH, U. G. (2008). Whole genome transcription profiling of Anaplasma phagocytophilum in human and tick host cells by tiling array analysis. BioMed Central Genomics. 9: 364. 120 NIELSEN, R. (2002). Mapping mutations on phylogenies. Syst. Biol. 51 , 729-739. NOH, S. M., BRAYTON, K. A , KNOWLES, D. P., AGNES, J. T. , DARK, M. J, BROWN, W, C., BASZLER, T. V. and PALMER, G. H. (2006). Differential expression and sequence conservation of the Anaplasma marginale msp2 gene superfamily outer membrane proteins. Infection and Immunity. 74: 3471-3479. NOH, S. M., BRAYTON, K. A , BROWN, W. C., NORIMINE, J., MUNSKE, G. R. , DAVITT, C. M. and PALMER, G. H. (2008). Composition of the surface proteome of Anaplasma marginale and its role in protective immunity induced by outer membrane immunization. Infection and Immunity. 76: 2219-2226. NOH, S. M., ZHUANG, Y., FUTSE, J. E., BROWN, W. C., BRAYTON, K. A and PALMER, G. H. (2010). The immunization-induced antibody response to the Anaplasma marginale Major Surface Protein 2 and its association with protective immunity. Vaccine. 28: 3741-3747. NORVAL, R. A I. , PERRY, B. D. and YOUNG, A S. (1992). The Epidemiology of theileriosis in Africa. Academic Press, London. NORVAL, R. A I. , SUTHERST, R. W. and KERR, J. D. (1996). Infestations of the bont tick Amblyomma hebraeum (Acari: lxodidae) on different breeds of cattle in Zimbabwe. Experimental and Applied Acarology. 20: 599-605. 121 NYANGIWE, N. and HORAK, I. G. (2007). Goats as alternative hosts of cattle ticks, Onderstepoort Journal of Veterinary Research. 74:1-7. OBERLE, S. M. and BARBET, A. F. (1993). Derivation of the complete msp4 gene sequence of Anaplasma marginale without cloning. Gene. 136: 291-294. OBERLE, S. M., PALMER, G. H. and BARBET, A. F. (1993). Expression and immune recognition of the conserved MSP4 outer membrane protein of Anaplasma marginale. Infection and Immunity. 61: 5245- 5251 . OCAIDO, M., MUWAZI, R. and OPUDA-ASIBO, J. (2009). Economic impact of ticks and tick-borne diseases on cattle production systems around Lake Mburo National Park in South Western Uganda. Tropical Animal Health and Production . 41: 731-739. OIE. (2012). Bovine Anaplasmosis. OIE Manual of Diagnostic Tests and Vaccines for Terrestrial Animals, seventh ed. Office International des Epizooties, Paris (Chapter 2.4.1) . PALMER, G. H. , KOCAN, K. M., BARRON, S. J., HAIR, J. A. , BARBET, A. F., DAVIS, W . C. and MCGUIRE, T. C. (1985). Presence of common antigens, including major surface protein epitopes, between the cattle (intraerythrocytic) and tick stages of Anaplasma marginale. Infection and Immunity. 50: 881-886. 122 PALMER, G. H., WAGHELA, S. D., BARBET, A. F., DAVIS, W. C. and MCGUIRE, T. C. (1987). Characterization of a neutralization-sensitive epitope on the Am 105 surface protein of Anaplasma marginale. International Journal of Parasitology. 17: 1279-1285. PALMER, G. H. (1989). Anaplasma vaccines, In I. G. Wright (ed.), Veterinary protozoan and hemoparasite vaccines. CRC Press, Boca Raton, Fla. p. 2-29. PALMER, G. H., EID, G., BARBET, A. F., MCGUIRE, T. C. and MCELWAIN, T. F. (1994). The immunoprotective Anaplasma marginale major surface protein 2 is encoded by a polymorphic multigene family. Infect. lmmun. 62:3808-3816. PALMER, G. H., RURANGIRWA, F. R., KOCAN, K. M. and BROWN, W. C. (1999). Molecular basis for vaccine development against the ehrlichial pathogen Anaplasma marginale. Parasitology Today. 15: 253-300. PALMER, G. H., RURANGIRWA, F. R. and MCELWAIN, T. F. (2001). Strain composition of the ehrlichia Anaplasma marginale with in persistently infected cattle, a mammalian reservoir for tick transmission. Journal of Clinical Microbiology. 39: 631- 635. PALMER, G. H., BANKHEAD, T. and LUKEHART, S. A. (2009). Nothing is permanent but change: Antigenic variation in persistent bacterial pathogens. Ce// Microbiology. 11: 1697-1705. PALMER, G. H. and BRAYTON, K.A. (2013). Antigenic variation and transmission fitness as drivers of bacterial strain structure. Ce//. Microbiology. 15: 1969-75. 123 PLAYFORD, M. (2005). Review of Research Needs for Cattle Tick Control Phases I and II Meat and Livestock Australia Limited, North Sydney, 1-162. POHL, A. E., CABEZAS-CRUZ, A. , RIBEIRO, M. F. B., SILVEIRA, J. A. G., SILAGHI , C., PFISTER, K. and PASSOS, L. M. F. (2013). Detection of genetic d iversity of Anaplasma marginale isolates in Minas Gerais, Brazil. Brazilian Journal of Veterinary Parasitology. 22: 129-135. POND, S. L. and FROST, S . D. (2005). Datamonkey: rapid detection of selective pressure on individual sites of codon alignments. Bioinformatics. 21: 2531-2533. PORTO NETO, L. R. , JONSSON, N. N., D'OCCHIO, M. J. and BARENDSE, W . (201 1). Molecular genetic approaches for identifying the basis of variation in resistance to tick infestation in cattle. Veterinary Parasitology. 180: 165-172. PUPKO, T., SHAMIR, I. P. R. and GRAUR, D. (2000). A fast algorithm for joint reconstruction of ancestral amino acid sequences. Molecular Biology and Evolution. 17: 890-896. REGASSA, A. , PENZHORN, B. L. and BRYSON, N. R. (2003). Atta inment of endemic stability to Babesia bigemina in cattle on a South African ranch where non-intensive tick control was applied. Veterinary Parasitology. 116: 267-27 4. 124 REINBOLD, J. B., COETZEE, J. F., SIRIGIREDDY, K. R. and GANTA, R. R. (2010). Detection of Anaplasma marginale and A. phagocytophilum in bovine peripheral blood samples by duplex real-time reverse transcriptase PCR assay. Journal of Clinical Microbiology. 48: 2424-2432. RIKHOTSO, B. 0 ., STOLT SZ, W. H., BRYSON, N. R. and SOMMERVILLE, J. E. (2005). The impact of 2 dipping systems on endemic stabil ity to bovine babesiosis and anaplasmosis in cattle in 4 communally grazed areas in Limpopo Province, South Africa. Journal of South African Veterinary Association. 76: 217-223. RODRIGUEZ, J. L. , PALMER, G. H., KNOWLES, D. P. and BRAYTON, K .A (2005). Distinctly different msp2 pseudogene repertoires in Anaplasma marginale strains that are capable of superinfection . Gene. 361: 127-132. RODRIGUEZ, S, D., ORTIZ, M.A.G., OCAMPO, R.J ., VEGA Y. and MURGUIA, C.A. (2009). Molecular epidemiology of bovine anaplasmosis with a particular focus in Mexico. Infection Genetics and evolution. 9: 1092-1101 . RUYBAL, P., MORETTA, R. , PEREZ, A , PETRIGH, R. , ZIMMER, P., ALCARAZ, E., ECHAIDE, I. , TORIONI DE ECHAIDE, S. , KOGAN, K. M., DE LA FUENTE, J. and FARBER, M. (2009). Genetic diversity of Anaplasma marginale in Argentina. Veterinary Parasitology. 162: 176- 180. 125 SACKETT, D., HOLMES, P., ABBOTT, K., JEPHCOTT, S. and BARBER, M. (2006). Assessing the Economic Cost of Endemic Disease on the Profitability of Australian Beef Cattle and Sheep Producers. Meat and Livestock Australia Limited, North Sydney, 1- 133. SAITOU, N. and NEI , M. (1987). The neighbor-joining method: a new method for reconstructing phylogenetic trees. Molecular Biology and Evolution . 4: 406-425. SANTOS, P. S., SENA, A A , NASCIMENTO, R. , ARAUJO, T. G., MENDES, M. M., MARTINS, J. R. , MINEO, T. W., MINEO, J. R. and GOULART, L. R. (2013). Epitope- based vaccines with the Anaplasma marginale Msp1a functional motif induce a balanced humoral and cellular immune response in mice. PLoS One. 8: e60311 . SCHNEIDER, A , CANNAROZZI , G. M. and GONNET, G. H. (2005). Empirical codon substitution matrix. BioMed Central Bioinformatics. 6: 134. SCHNEIDER, A , GONNET, G. and CANNAROZZI , G. (2007). SynPAM - a distance measure based on synonymous codon substitutions. IEEE/ACM Trans. Computational Biology Bioinformatics. 4: 553- 560. SCOLES, G. A , BROCE, A B., LY SYK, T. J. and PALMER, G. H. (2005). Relative efficiency of biological transmission of Anaplasma marginale (Rickettsiales: Anaplasmataceae) by Dermacentor andersoni (Acari: lxodidae) compared with mechanical transmission by Stomoxys calcitrans (Diptera: Muscidae). Journal of Medical Entomology. 42: 668- 675. 126 SHANNON, C.E. (1948). The mathematical theory of communication. Bell System Technical Journal. 27: 379-423. SHKAP, V., BRAYTON, K., BROWN, W. C. and PALMER, G. H. (2002). Expression of major surface protein 2 variants with conserved T-cell epitopes in Anaplasma centrale vaccinates. Infection and Immunity. 70: 642-648. SILVESTRE, B. T. , RABELO, E. M., VERSIANI, A. F., DA FONSECA, F. G., SILVEIRA, J. A. , BUENO, L. L. , FUJIWARA, R. T. and RIBEIRO, M. F. (2014). Evaluation of humoral and cellular immune response of BALB/c mice immunized with a recombinant fragment of Msp1a from Anaplasma marginale using carbon nanotubes as a carrier molecule. Vaccine, http://dx.doi.org/10.1016/j .vaccine.2014.02.062 (Epub ahead of print). STEVENS, K. B., SPICKETT, A. M., VOSLOO, W., PFEIFFER, D. U., DYASON, E. and DU PLESSIS, B. (2007). Influence of dipping practices on the seroprevalence of babesiosis and anaplasmosis in the foot-and-mouth disease buffer zone adjoining the Kruger National Park in South Africa. Onderstepoort Journal of Veterinary Research. 74: 87-95. STIK, N. I. , ALLEMAN, A. R. , BARBET, A. F., SORENSON, H. L., WANSLEY, H. L. , GASCHEN, F. P., LUCKSCHANDER, N., WONG, S., CHU, F., FOLEY, J. E., BJOERSDORFF, A. , STUEN, S. and KNOWLES, D.P. (2007). Characterization of Anaplasma phagocytophilum major surface protein 5 and the extent of its cross- reactivity with A. marginale. Clinical Vaccine Immunology. 14: 262-268. 127 TAMURA, K. , DUDLEY, J., NEI, M. and KUMAR, S. (2007). MEGA4: Molecular Evolutionary Genetics Analysis (MEGA) software version 4.0. Molecular Biology and Evolution. 24: 1596-1599. TAMURA, K. , PETERSON, D., PETERSON, N., STECHER, G., NEI , M. and KUMAR, S. (2011 ). MEGA5: Molecular Evolutionary Genetics Analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution. 28(10): 2731-2739. TEBELE, N., MCGUIRE, T. C. and PALMER, G. H. (1991). Induction of protective immunity by using Anaplasma marginale initial body membranes. Infection and Immunity. 59: 3199-3204. THEILER, A. (1910). Anaplasma marginale (gen and spec., nov.). The marginal points in the blood of cattle suffering from a specific disease. Report of the Government on Veterinary Bacteriology in Transvaal , Department of Agriculture 1908-1909. THEILER, A. (1911 ). Further investigations into anaplasmosis of South African cattle. In: 1st Report of the Director of Veterinary Research. Department of Agriculture of the Union of South Africa, 7-46. THOMPSON, J. D., HIGGINS, D. G. and GIBSON, T. J. (1994). CLUSTAL W : improving the sensitivity of progressive multiple sequence alignment through sequence weighting, positions-specific gap penalties and weight matrix choice. Nucleic. Acids Research. 22: 4673-4680. 128 TORINA, A. , MORENO-CID, J. A. , BLANDA, V ., FERNANDEZ DE MERA, I. G., DE LA LASTRA, J. M., SCIMECA, S., BLANDA, M., SCARIANO, M. E., BRIGAN6, S., DISCLAFANI, R. , PIAZZA, A. , VICENTE, J., GORTAZAR, C., CARACAPPA, S., LELLI , R.C. and DE LA FUENTE, J. (2014). Control of tick infestations and pathogen prevalence in cattle and sheep farms vaccinated with the recombinant Subolesin-Major Surface Protein 1a chimeric antigen. Parasitology Vectors Biology. 7: 10. TORIONI DE ECHAIDE, S., KNOWLES, D. P., MCGUIRE, T. C., PALMER, G. H., SUAREZ, C. E. and MCELWAIN, F. F. (1998). Detection of cattle naturally infected with Anaplasma marginale in a reg ion of endemicity by nested PCR and a competitive enzyme-linked immunosorbent assay using recombinant major surface protein 5. Journal of Clinical Microbiology. 36: 777-782. TURTON J. A. , KATSANDE T. C., MATINGO M. B., JORGENSEN W. K., USHEWOKUNZE-OBATOLU U. and DALGLIESH R. J. (1998). Observations on the use of Anaplasma centrale for immunization of cattle against anaplasmosis in Zimbabwe. Onderstepoort Journal of Veterinary .65:81- 86. TURTON, J. (2001 ). External parasites of cattle. Department of Agriculture, S.A. , http://www.nda.agric.za/docs/parasites/parasites.htm, 18 May 2013. VIDOTTO, M. C., KANO, S. F., GREGORI, F. , HEADLEY, S. A. and VIDOTTO, 0 . (2006). Phylogenetic analysis of Anaplasma marginale strains from Parana State, Brazil, using the msp1 a and msp4 genes. Journal Veterinary Medicine. 53: 404-411 . 129 VISESHAKUL, N., KAMPER, S., BOWIE, M. V. and BARBET, A F. (2000). Sequence and expression analysis of a surface antigen gene family of the rickettsia Anaplasma marginale. Gene. 253: 45-53. VISSER, E. S., MCGUIRE, T. C., PALMER, G. H. , DAVIS, W. C., SHKAP, V., PIPANO, E. and KNOWLES, D.P. (1992). The Anaplasma marginale msp5 gene encodes a 19- kilodalton protein conserved in all recognized Anaplasma species. Infection and Immunity. 60: 5139-5144. WALKER, J. B. 1991 . A review of the ixodid ticks (Acari , lxodidae) occurring in southern Africa. Onderstepoort Journal of Veterinary Research. 58: 81-105. WAMBURA, P. N., GWAKISA P. S., SILAYO R. S. and RUGAIMUKAMU E. A (1998). Breed-associated resistance to tick infestation in Bos indicus and their crosses with Bos taurus. Veterinary Parasitology. 77: 63-70. WARNER, C. K. and DAWSON, J. E. (1996). Genus- and species-level identification of Ehrlichia species by PCR and sequencing. In PCR protocols for emerging infectious diseases. Edited by Persing DH. Washington DC: ASM Press; 100-105. WEN, B., JIAN, R. , ZHANG, Y. and CHEN, R. (2002): Simultaneous detection of Anaplasma marginale and a new Ehrlichia species closely re lated to Ehrlichia chaffeensis by sequence analysis of 16S ribosomal DNA in Boophilus microplus ticks from Tibet. Journal of Clinical Microbiology. 40: 3286- 3290. 130 YANG, Z., KUMAR, S. and NEI , M. (1995). A new method of inference of ancestral nucleotide and amino acid sequences. Genetics. 14: 1641-1650. YU, X. J., ZHANG, X. F., MCBRIDE, J. W., ZHANG, Y. and WALKER, D. H. (2001 ).Phylogenetic relationships of Anaplasma marginale and 'Ehrlichia platys' to other Ehrlichia species determined by GroEL aa sequences. International Journal of System Evolution Microbiology. 51(3): 1143-1146. ZHANG, Z., SCHWARTZ, S., WAGNER, L. and MILLER, W. (2000). A greedy algorithm for aligning DNA sequences. J. Comput. Biol. 7: 203-214. ZIAM, H. and BENAOUF, H. (2004). Prevalence of blood parasites in cattle from wilayates of Annaba and El Tarf east Algeria. Arch. Inst. Pasteur Tunis. 81: 27-30. ZIVKOVIC, Z. (2010). Tick-pathogen interactions in bovine anaplasmosis. Doctoral Thesis, University of Utrecht, Netherland. 131 APPENDICES Appendix1: Raw data of statistical analysis of A marginale in cattle in studied provinces of South Africa The table below qives the qeneral outcome . tab Anapla sma_results nw ANA PLASMA I MARGINALE I PCR RESULTS I Freq . Percent Cum . - -----------+----------- ------------- ----------- Negative I 71 25 . 36 Positive I 209 100 . 00 - - -- - - - -----+------- - ---- - ---------------------- Total I 280 100 . 00 . prop Anaplasma_ results_ nw Proportion estimation Number of obs 280 I Pr oportion Std . Err . [95% Conf . Interval) --- ------------------+---------------- -------------- ------------------ Anaplasma_ resul ts nw I Negative I . 0260461 . 2057817 . 308153 4 Positive I . 0260461 . 6918 466 . 7942183 . tab Anaplasma_results nw farm_nw , exact ANAPLASMA MARGI NALE PCR I FARM RESULTS I Commer cia Communal I Total -----------+----------------------+---------- Negati ve I 4 67 I 71 Pos it i ve I 61 148 I 209 ---------- -+----------------------+--- - - ----- Total I 65 215 I 280 Fi sher ' s exact 1-si ded Fisher ' s exact 0 . 000 132 . prop Anaplasma_results_nw , over( farm_nw) Proportion estimation Number of obs 280 Negative: Anaplasma_resul t s_nw Negative Positive : Anaplasma_results_nw Positive Commercial : farm nw =Commercial Communal : farm nw = Communal Over I Proportion Std . Err. [ 95% Conf. Interva l) -------------+------------------------------------------------ Negative Commercial I . 0615385 . 030039 4 . 0230112 . 1543787 Communa l I . 3116279 . 0316609 . 2529283 . 3770745 -------------+--------------------------------------------- --- Positive Commercial I . 0300394 . 84 56213 . 9769888 Communal I . 0316609 . 6229255 . 7470717 . kwallis Anaplasma_results_nw , by( location_nw) Kruskal-Wallis equality-o f-populations rank test +--------------------------------+ location_nw I Obs I Rank Sum I !---------------+-----+----------! I Eastern Cape I 40 I 7040 . 00 I I Gauteng I 20 I 3520 . 00 I I KwaZulu -Natal I 55 I 8700 . 00 I Limpopo I 40 I 5920 . 00 I I Mpumalanga I 21 I 3696 . 00 I !---------------+-----+----------! I North West I 33 I 4828 . 00 I I Northern Cape I 45 I 1620 . 00 I I Western Cape I 26 I 4016 . 00 I +------------------- -------------+ chi- squared 94 . 425 with 7 d . f . probability 0 . 000 1 133 The hypothesis that the farms are similar with respect to prevalence is rejected by the results above . tab location_nw Anaplasma_results_nw , row col cell chi2 +-------------------+ I Key I 1-------------------1 I frequency I I row percentage I column percentage I cell percentage I +-------------------+ I ANAPLASMA MARGINALE I PCR RESULTS LOCATION I Negative Positive Total --------------+----------------------+---------- Eastern Cape I O 40 I 40 o. oo 100 . 00 I 100 . 00 0 . 00 I 14 . 29 0 . 00 I 14.29 111111 --------------+----------------------+---------- Gauteng I 0 20 I 20 I 0 . 00 100 . 00 I 100 . 00 I 0 . 00 .. I 7 . 14 I o. oo 7 . 14 I 7 . 14 --------------+----------------------+---------- KwaZulu-Natal I 7 48 I 55 I 12 . 73 87 . 27 I 100 . 00 I 9 . 86 I 19 . 64 I 2 . 50 I 19 . 64 --------------+--- -------------------+---------- Limpopo I 8 32 I 40 I 20 . 00 80 . 00 I 100 . 00 I 11 . 27 111111 I 14 . 29 I 2 . 86 11 . 43 I 14 . 29 --------------+--- -------------------+---------- Mpumalanga I 0 21 I 21 I 0 . 00 100 . 00 I 100 . 00 I 0 . 00 I 7 . 50 I 0 . 00 I 7 . 50 --------------+----------------------+---------- North West I 7 26 I 33 I 21. 21 78 . 79 I 100 . 00 I 9 . 86 111111 I 11 . 79 I 2 . 50 9 . 29 I 11 . 79 --------------+----------------------+---------- Northern Cape I 4 5 0 I 4 5 I 100 . 00 o. oo I 100 . 00 I 63 . 38 o. oo I 16 . 07 I 16 . 07 .. I 16 . 07 Lowest (Northern Cape) --------------+----------------------+---------- 134 Western Cape I 4 22 I 26 15 . 38 8 4 . 62 I 100 . 00 5 . 63 - I 9 . 29 1. 43 7 . 86 I 9.29 --------------+----------------------+------ ---- Total I 71 209 I 280 I 25 . 36 74 . 64 I 100 . 00 I 1 00 . 00 100 . 00 I 100 . 00 I 25 . 36 74 . 64 I 100.00 Pearson chi2(7) = 1 66 . 8890 Pr 0 . 000 . prop Anaplasma results nw if farm nw==lllllllllllll Proport i on estimation Number of obs 65 I Proportion Std . Er r . [95% Cont . Interval] -- - - --------- -- ------+-- --- ------------------------------- ------------ Anaplasma_ results_ nw I Negative I . 0615385 . 0300394 . 0226719 . 1563737 Positive I . 938461 5 . 0300394 . 843 6263 . 977 32 81 ~aplasm Proportion estimation Number o f obs 215 I Pr oportion St d . Err. [95% Conf. Inte r val ] - - --------------- ----+----------------- ----------------------------- - - Anaplasma_ result s nw I Negat i v e I . 3116279 . 0316609 .2528555 . 377 165 Positive I . 6883721 . 0316609 . 622835 . 7471445 tab farm nw FARM I Freq . Percent Cum . ----- -------+----------------------------------- Commercial I 65 23 . 2 1 23 . 21 Communal I 215 76 . 79 100 . 00 ----- - ------+----- ------------ ------------ - ----- Total I 280 100 . 00 . kwallis location_ nw if Anaplasma_results_nw==O , by( farm nw) Kruskal-Wallis equality-of-populations rank test +------------- ------ - ---------+ I farm nw I Obs I Rank Sum I 135 1------------+- -- - -+----------1 I Commercial I 4 I 27 8 . 00 I I Communal I 67 I 2 278 . 00 I +-- - ----- - - -------- - - - - -------+ c hi-squared 11 . 167 wi th 1 d . f . p r obability 0 . 000 8 chi - squared with ties = 15 . 049 wi t h 1 d . f . probabil i t y 0 . 000 1 . kwalli s loc ation_nw i f Anaplasma_ r esults_lllllllllll) , by( farm nw) Kruskal - Wall is equa lity- o f - population s rank t es t +----- - -- - - - - --- - -------------+ farm_nw I Obs I Ran k Sum I !--- - - ------- +--- - -+- ------- - - ! I Commerc ial I 61 I 5 7 66 . 50 I I Communal I 148 I 1 61 78 . 50 I +--------------- - --- - ---- ---- -+ chi- square d 2 . 580 with 1 d . f . p r obabi l i t y 0 . 1082 c h i-squared with t ies = 2 . 654 with 1 d . f . prob a bility 0 . 1033 . tab Anaplasma_results nw ANAPLASMA I MARGINALE I PCR RESULTS I Fr e q . Per c ent Cum. - - ----------+-- - - --- ---------- - -- - ----------- -- - Negativ e I 71 2 5 . 3 6 2 5 . 3 6 Po s itive I 209 74. 64 1 0 0 . 00 - - - - -- - - - ---+- - --- --- - ---- - - -- - ----- - ------ ----- To tal I 280 100 . 00 . p r op location_ nw if Anaplasma_results_nw== l~) , ove r ( f a rm nw) Propo rtio n estimatio n Number o f obs 209 Over I Prop o rtio n Std . Err . [ 95% Conf. Interv al ) - ------ - - ----+- ----- - --- --- - ------ - -- - ----------- - - - ---- ------ Eastern Cape I Commerc i al I . 31147 5 4 . 05978 55 . 2070 474 . 4393882 136 Communal I .14 18919 . 02878 . 0940079 . 2085517 -------------+------------------------------------------------ Gauteng Commercial I . 3278689 . 060604 . 2209735 . 4561933 Communal I . (no observations) -------------+------------------------------------------------ KwaZulu-Natal I Commercial I (no observations) Communal I . 3243243 . 03861 . 2532568 .4 0 4531 -------------+------------------------------------------------ Limpopo Commercial I (no observat ions ) Communal I . 2162162 . 0339534 . 1567198 . 2905182 -------------+------------------------------------------------ Mpumalanga I Commercial I (no obse r vatio ns) Communal I .14 18919 . 02878 . 09 40079 . 2085517 -------------+------------------------------------------------ North West I Commercial I (no observations) Communal I .1 756757 . 0313867 . 1220444 . 2 462639 -------------+------------------------------------------------ Western Cape I Commercial I . 3606557 . 061992 4 . 2 4 92643 .4 89377 Communal I . (no observ ations) . prop location_nw if Anaplasma_results nw==l & farm nw==l , over( farm_nw) Proportion estimation Number of obs 61 Commercial : farm nw - Over I Proportion Std. Err . (95 % Conf . Interval) -------------+-------------------------- ---------------------- Eastern Cape I Commercial I .3114 7 54 . 0597855 . 2057294 . 4413712 -------------+-------------------------- ---------------------- Gauteng I Commerc ial I • 3278689 . 06060 4 . 2 1961 01 . 4581632 -- -----------+----------- ------------------------------------- Western Cape I Commercial I . 3606557 . 0619924 . 24 7815 . 4913163 . prop l ocation_nw if Anaplasma_results_ nw==l & , over( farm_nw) Over I Prop ortion Std . Err . (95% Conf. I nterval) 137 -------------+------------------------------------------------ Eastern Cape I Communal I . 1418919 . 02878 . 0939113 . 208739 -------------+------------------------------------------------ KwaZulu - Natal 1 Communal I . 3243243 . 03861 . 253097 . 4047346 -------------+------------------------------------------------ Limpopo Communal I . 2162162 . 0339534 . 1565927 . 2907164 -------------+------------------------------------------------ Mpumalanga Communal I . 1418919 . 02878 . 0939113 . 2087 39 -------------+------------------------------------------------ North West I Communal I . 1756757 . 0313867 . 121933 . 2464571 . prop location_nw if Anaplasma results_nw==O , over ( farm nw) Proporlion estimation Number of obs 71 Commercial : farm nw Commercial Communal : farm nw =Communal Over I Proportion Std . Err . [95% Conf . Interva l) -------------+------------------------------------------------ KwaZulu - Natal I Commercial I . (no observations ) Communal I . 1044776 . 0376511 . 0496877 . 206552 -------------+------------------------------------------------ Limpopo Commercial I . (no observations) Communal I . 119403 . 0399139 . 059794 . 2242624 -------------+------------------------------------------------ North West I Commercial I . (no observat i ons) Communa l I . 10 4 4776 . 0376511 . 0496877 . 206552 -------------+--------- - -------------------------------------- Northern Capel Commercial I . (no observations) Communal I . 6716418 . 0578057 . 5480652 . 7752826 -------------+-------------- - --------------------------------- Western Cape I Commercial I 1 0 Communal I . (no observations) 138 Appendix 2: GenBank accession number of Anaplasma marginale msp1a gene isolates and their origin in South Africa Sample names Origin Gen Bank accession numbers LP-37 Limpopo (KC470153) LP-46 Limpopo (KC470154) LP-7 Limpopo (KC470155) LP-10 Limpopo (KC470156) LP-50 Limpopo (KC470157) LP-30 Limpopo (KC470158) LP-34 Limpopo (KC470159) MP-C2 Mpumalanga (KC4701 60 MP-C5 Mpumalanga (KC470161) NW-C2 North West (KC4701 62) NW-C4 North West (KC470163) NW-C5 North West (KC4701 64) NW-C1-160312 North West (KC470165) NW-C4-160312 North West (KC470166) GP-C1 Gauteng (KC470167) GP-C2 Gauteng (KC470168) GP-C5 Gauteng (KC470169) 139 GP-C1112105 Gauteng (KC470170) GP-C4117105 Gauteng (KC470171 ) GP-C1817105 Gauteng (KC470172) GP-C7117105 Gauteng (KC470173) KZN-F KwaZulu-Natal (KC470174) KZN-MN KwaZulu-Natal (KC470175) KZN-K KwaZulu-Natal (KC470176) KZN-D KwaZulu-Natal (KC470177) KZN-Y KwaZulu-Natal (KC470178) KZN-14 KwaZulu-Natal (KC470179) KZN-19 KwaZulu-Natal (KC470180) KZN-49 KwaZulu-Natal (KC470181 ) KZN-51 KwaZulu-Natal (KC470182) EC-22 Eastern Cape (KC470183) EC-23 Eastern Cape (KC470184) EC-24 Eastern Cape (KC470185) WC-4 Western Cape (KC470186) WC-6 Western Cape (KC470187) WC-7 Western Cape (KC470188) WC-8 Western Cape (KC470189) WC-10 Western Cape (KC470190) 140 WC-11 Western Cape (KC470191) WC-12 Western Cape (KC470192) WC-13 Western Cape (KC470193) WC-14 Western Cape (KC470194) WC-15 Western Cape (KC470195) WC-16 Western Cape (KC470196) 141 Appendix 3: GenBank accession number of Anap/asma marginale msp4 gene isolates and their origin in South Africa Samples name Origin Gen Bank accession number LP-C1 Limpopo Province KF758907 LP-C2 Limpopo Province KF758909 LP-C3 Limpopo Province KF758910 LP-C7 Limpopo Province KF758911 LP-C9 Limpopo Province KF758912 LP-C10 Limpopo Province KF758913 LP-C11 Limpopo Province KF758914 LP-C14 Limpopo Province KF758915 LP-C19 Limpopo Province KF758916 LP-C21 Limpopo Province KF758917 LP-C23 Limpopo Province KF758918 LP-C27 Limpopo Province KF758919 LP-C29 Limpopo Province KF758920 LP-C30 Limpopo Province KF758921 LP-C32 Limpopo Province KF758922 LP-C33 Limpopo Province KF758923 LP-C34 Limpopo Province KF758924 142 LP-C35 Limpopo Province KF758925 LP-C38 Limpopo Province KF758926 LP-C40 Limpopo Province KF758927 LP-C41 Limpopo Province KF758928 LP-C43 Limpopo Province KF758929 LP-C46 Limpopo Province KF758930 LP-C49 Limpopo Province KF758931 LP-C50 Limpopo Province KF758932 LP-C77 Limpopo Province KF758933 LP-C85 Limpopo Province KF758934 MP-C1 Mpumalanga Province KF758860 MP-CS Mpumalanga Province KF758862 MP-C? Mpumalanga Province KF758864 MP-C8 Mpumalanga Province KF758866 MP-C15 Mpumalanga Province KF758867 MP-C18 Mpumalanga Province KF758869 MP-C19 Mpumalanga Province KF758871 MP-C20 Mpumalanga Province KF758872 MP-C21 Mpumalanga Province KF758874 NW-CB3 North West Province KF758842 NW-CB4 North West Province KF758876 143 NW-CBS North West Province KF7S8878 NW-CB? North West Province KF7S8880 NW-CB8 North West Province KF7S8882 NW-C28 North West Province KF7S8883 GP-C33 Gauteng Province KF7S8847 GP-C48 Gauteng Province KF7S8849 GP-CS6 Gauteng Province KF7S88S2 KZN-C26 KwaZulu-Natal Province KF7S88S4 KZN-C29 KwaZulu-Natal Province KF7S88S6 KZN-C31 KwaZulu-Natal Province KF7S88S7 KZN-C37 KwaZulu-Natal Province KF7S88S8 KZN-C42 KwaZulu-Natal Province KF7S88S9 EC-C6A Eastern Cape Province KF7S8843 EC-C2SA Eastern Cape Province KF7S884S EC-C33A Eastern Cape Province KF7S8847 WC-C4 Western Cape Province KF7S888S WC-CS Western Cape Province KF7S8886 WC-C6 Western Cape Province KF7S8888 WC-C7 Western Cape Province KF7S8890 WC-C8 Western Cape Province KF7S8892 WC-C9 Western Cape Province KF7S8894 144 WC-C10 Western Cape Province KF758896 WC-C12 Western Cape Province KF758898 WC-C15 Western Cape Province KF758900 WC-C16 Western Cape Province KF758902 WC-C18 Western Cape Province KF758904 WC-C20 Western Cape Province KF758905 145 r I f Appendix 4: Amplified sequences of Anaplasma spp. isolates , their origin and GenBank accession number Isolates and their origin GenBank accession number California AY010248.1 GZ8-China JN572928.1 51 (G18)-Hungary EU315783.1 New Castle AY127077.1 Glencoe 2 AY127068.1 Stillwater 1 AY127075.1 AY851150.1 Switzerland Oregon AY127065.1 661 Kari-Kenya AY666004.1 USA AF428081 .1 1.6-Zimbabwe AY666006.1 2.4-Zimbabwe AY666007.1 5.9-Zimbabwe AY666011 .1 Italy 6 AY702917.1 Italy 47 AY702921 .1 Israeli non-tailed AY786993.1 50 (G16)-Hungary EU315782.1 Mississippi AY010251 .1 146 Virg inia AY010254.1 Oklahoma City AY127073.1 Israeli tailed AY786994.1 Okeechobee AY010253.1 Stillwater 2 AY127076.1 Israeli round AY787172.1 SP?-Spain AY456002.1 TWN1-Taiwan EU677383.1 Mexico AF428083.1 Mexico AF428084.1 Mexico AF428085.1 Brazil 4 AY283190.1 Brazil 5 AY283191 .1 Puerto Rico AY191827.1 F72-Austral ia AY666002.1 G38-Austra lia AY666003.1 Argentina AF428086.1 Anaplasma phagocytophilum AY706389 Anaplasma ovis AF393742.1 147